Site Investigation, Clayton, Matthews and Simons.pdf

January 21, 2018 | Author: adeeb | Category: Geotechnical Engineering, Civil Engineering, Engineering, Geology, Science
Share Embed Donate


Short Description

Download Site Investigation, Clayton, Matthews and Simons.pdf...

Description

Site Investigation

Second Edition

C. R. I. Clayton, M. C. Matthews and N. E. Simons Department of Civil Engineering, University of Surrey

Contents Preface 1 Planning and procurement Introduction Objectives General design philosophy Implementation Planning ground investigations Procurement Execution 2 Description and classification of soils and rocks Introduction Soil and rock description Soil description Soil classification Rock description Description of rock material Description of discontinuities Methods of collecting discontinuity data Discontinuity surveys Presentation of discontinuity data Description of rock masses Records of boreholes State of recovery of core Records of trial pits and shafts 3 The desk study and walk-over survey Introduction Sources of information for desk studies Air photography and remote sensing Satellite remote sensing The walk-over survey 4 Subsurface exploration: engineering geophysics Introduction Lateral variability Profiling Sectioning Determination of properties 5 Subsurface exploration: boring, drilling, probing and trial pitting Introduction Boring Drilling Probing Examination in situ 6 Sampling and sample disturbance Introduction Sample sizes Soil disturbance Classification of soil samples British practice, and the BS 5930 classification 7 Undisturbed sampling techniques Introduction

ii

Contents Samples from pits and exposures Drive samplers Rotary samplers Sand sampling Sampler selection 8 Laboratory testing Introduction The purpose of soil testing Available tests Consolidation tests Accuracy and measuring systems 9 In situ testing Introduction Penetration testing Strength and compressibility testing Permeability testing 10 Basic field instrumentation for site investigation Introduction Uses of instrumentation Requirements for instrumentation Pore water pressure and groundwater level measurement Displacement measurement Other measurements References and standards Index

iii

Preface It is now over 12 years since we completed the first edition of this book. In the intervening time there have been a number of important advances in the way that site investigations are carried out, both in the UK and elsewhere. In this new edition we have described those techniques which are now in regular commercial use, some of which were presaged in our first edition. But, as in the first edition of this book, we have avoided including descriptions of techniques which we believe will remain largely in the research field. For the second edition we have added substantial new material on: • • • • • •

specification and procurement; desk studies; geophysical investigation techniques; sample disturbance and sampling methods; in situ testing; and laboratory testing.

The object of this book remains the same: we aim to improve the quality of site investigation by providing a relatively simple and concise reference book intended to be read by civil and structural engineers and engineering geologists, when undergraduates and postgraduates, but particularly when in practice. The text is intended to inform the reader of available techniques, and to illustrate the advantages and disadvantages of these techniques. Site investigation is a complex process. It is vital to the success of any construction project, since inadequate investigation can lead to very large construction cost overruns. If site investigation is to be effective then it must be carried out in a systematic way, using techniques that are relevant, reliable and cost-effective. We hope that readers will find this text a useful introduction to this important part of the building and construction process.

4

Chapter 1

Planning and procurement Unfortunately, soils are made by nature and not by man, and the products of nature are always complex. Karl von Terzaghi, 1936

INTRODUCTION

Site investigation is the process by which geological, geotechnical, and other relevant information which might affect the construction or performance of a civil engineering or building project is acquired. Soil and rock are created by many processes out of a wide variety of materials. Because deposition is irregular, soils and rocks are notoriously variable, and often have properties which are undesirable from the point of view of a proposed structure. Unfortunately, the decision to develop a particular site cannot often be made on the basis of its complete suitability from the engineering viewpoint; geotechnical problems therefore occur and require geotechnical parameters for their solution. Site investigation will often be carried out by specialists in the field of soil mechanics. Soil, in the engineering sense, is the relatively soft and uncemented material which overlies the rock of the outer part of the Earth’s crust. Specialists in the mechanical behaviour of soil are normally civil engineers and in the UK they will often have some postgraduate geotechnical education: such people are termed ‘soils engineers’ or ‘geotechnical engineers’. Geologists with an interest in the relevance of geology to civil engineering or building construction are called ‘engineering geologists’. Soil mechanics in its present form is a relatively recent addition to the field of engineering. Interest in the behaviour of earth and rock for engineering purposes can be traced back to Roman times (Palladius in Plommer (1973)), but significant advances in analysis seem to date back to the eighteenth century, when the need for large defensive revetments led to early work on retaining walls. Coulomb’s paper, delivered to the Académie Royale des Sciences in 1773 and published in 1776, represents an early work which showed considerable understanding, inter alia, of the behaviour of soil, and whose results are still valid and in use (Heyman 1972). Subsequent papers, principally delivered by the French, did much to refine the available solutions but little to increase fundamental knowledge. By the first quarter of the nineteenth century, it appears that many concepts now associated with the principle of effective stress were intuitively understood. Telford used pre-loading during the construction of the Caledonian Canal in 1809 ‘for the purpose of squeezing out the water and consolidating the mud’, and Stephenson used drains to lower pore pressures during the construction of the Chat Moss embankment on the Liverpool and Manchester Railway in the years 1826 to 1829 ‘in order to consolidate the ground between them on which the road was to be formed’ (Smiles 1874). During the industrial period preceding the twentieth century, many of the currently used geotechnical processes for the improvement of ground, such as piling, pre-loading, compaction and de-watering appear to have been used (Feld 1948; Skempton 1960b; Jensen 1969). These techniques were applied in a purely empirical manner. At the turn of the twentieth century, a series of major failures occurred which led to the almost

1

Planning and Procurement simultaneous formation of geotechnical research groups in various countries. In America, slope failures on the Panama Canal led to the formation of the American Foundations Committee of the American Society of Civil Engineers in 1913 and, in Sweden, landslides during a railway construction resulted in the formation of the State Geotechnical Commission in the same year. Following a number of embankment and dyke failures, a government committee under Buisman was set up in Holland in 1920. Casagrande (1960), however, dates the advent of modern soil mechanics to the period between 1921 and 1925, when Terzaghi published several important papers relating to the pore pressures set up in clay during loading, and their dissipation during consolidation, and also published his book Erdbaumechanik auf Bodenphysikalischer Grundlage. These works largely stemmed from Terzaghi’s appreciation of the need to supplement geological information with numerical data, following two years spent collecting geological information on the construction sites of US dams (Terzaghi 1936). Terzaghi’s first professional work in England was in 1939, when he was retained to investigate a slope failure at the Chingford reservoir (Cooling and Golder 1942). As a result, the first commercial soil mechanics laboratory in the UK was established by John Mowlem and became Soil Mechanics Ltd in 1943. Whyte (1976) reports that by 1948 five other contractors and one consultant had soils divisions. Major encouragement was given to soils research in the UK by Cooling, who influenced a number of engineers (for example Skempton, Bishop and Golder) who worked at the Building Research Station in the 1940s. In 1948, Géotechnique commenced publication, and by 1955 a great number of significant papers on soil mechanics had been published covering topics such as site investigation, seepage, slope stability and settlement. According to Mayniel (1808), Bullet was the first to try to establish an earth pressure theory, in 1691. More importantly from our point of view, Bullet notes the importance of site investigation for the foundations of earth-retaining structures and recommends the use of trial holes in order to determine the different beds of soil beneath a site, and in order to ensure that poor soil does not underlie good soil. Where trial holes could not be made, Bullet recommended the use of an indirect method of investigation whereby the quality of the soil was determined from the sound and penetration achieved when it was beaten with a 6—8 ft length of rafter. Whilst the use of trial holes to investigate sub-soil may, not unexpectedly, date from centuries ago, it is more surprising to note that the equipment for boring holes in soft ground also has a long history. Jensen (1969) and Whyte (1976) illustrate types of drilling equipment in use around 1700, and many of the tools bear a striking resemblance to those used in light percussion drilling at the present time in the UK. Modern site investigation differs from its forbears principally because of the need to quantify soil behaviour. Terzaghi, in his James Forrest lecture to the Institution of Civil Engineers in London (1939) noted that in 1925 sampling methods in the USA were ‘primitive’, with sealed tube samples being almost unheard of. The work of Casagrande between 1925 and 1936 demonstrated the influence of soil disturbance during sampling (see, for example, Casagrande (1932)) and led to the development in the USA of ‘elaborate and ingenious procedures for furnishing almost undisturbed samples up to a diameter of 5-inches’ (Terzaghi 1939). At the same time considerable advances were made in Denmark, France, Germany, Sweden and England. In the UK, Cooling and Smith (1936) reported an early attempt at the acquisition of ‘undisturbed’ soil samples using a 105 mm dia. split tube forced into the ground from the back of a lorry. By 1937 the tool was a 105 mm dia. tube which was driven into the soil (Cooling and Golder 1942; Cooling 1942), and which had an area ratio (the ratio of displaced soil area to sample area) of about 20%. Boring was by well-boring apparatus, ‘sunk in the usual way with augers, chisels, etc.’ (Cooling 1942). By 1945 the sampling tube had become the U100 which is still in use today (Longsdon 1945). In 1949, the first draft Civil Engineering Code of Practice for Site Investigations was issued for

2

Site Investigation comment. At that time Harding (1949) delivered a paper to the Works Construction Division of the Institution of Civil Engineers in which he detailed the methods of boring and sampling then available. The recommendations made in that paper, and in discussions on the paper by Skempton, Toms and Rodin form the basis of the majority of techniques still in use in site investigation in the United Kingdom. For example, in his discussion on methods of boring, Harding notes that: the boring equipment used in site investigations is criticized by some who have not been exposed to the need to carry it themselves, as being primitive and lacking in mechanization. Whilst it is possible to think of many ingenious contrivances for removing articles at depths below ground, in practice simple methods usually prove to be more reliable.

while Skempton confirmed this view: with that simple equipment [shell and auger gear and 102 mm dia. sampler] the majority of site investigations in soils could be carried out and, moreover, sufficient experience was now available to enable the positive statement to be made that, in most cases, the results obtained by that technique (in association with laboratory tests) were sufficiently reliable for practical engineering purposes.

By 1953, Terzaghi stated in connection with site investigation that ‘we have acquired all the knowledge which is needed for a rational interpretation of the observational and experimental data’. The reader may reasonably ask what is to be gained from this book, since techniques are so well established. In reality, since 1950, four main changes have taken place. First, many of the methods introduced before and since have been the object of criticism as a result of differences between predictions and subsequent observations. Secondly, a considerable number of the lessons learnt before 1950 have been forgotten: few U100 samplers in use today are of the standard required by Hvorslev (1949) for undisturbed sampling, and much fieldwork remains unsupervised by engineers. Thirdly, few engineers have an experience or understanding of the techniques of boring and drilling holes for site investigations, and most clients remain unaware of the importance of this part of the work. Finally, recent years have seen the introduction of sophisticated and expensive methods of testing and computer analysis which cannot be sensibly applied to samples and predictions of soil conditions of indeterminate quality. The Civil Engineering Code of Practice No. 1: Site Investigations was issued in 1950, and revised as British Standard Code of Practice CP 2001 in 1957. This code has now been extended, completely rewritten and re-issued as British Standard 5930:1 981. At the time of writing (1992) BS 5930 is under revision. The code contains much valuable information, but it is perhaps necessary to ask whether it is wise to codify in this way. Terzaghi (1951) argued that: since there is an infinite variety of subsoil patterns and conditions of saturation, the use of the different methods of subsoil exploration cannot be standardised, but the methods themselves still leave a wide margin for improvement, as far as expediency and reliability are concerned.

OBJECTIVES

The objectives of site investigation have been defined by the various Codes of Practice (BS CP 2001:1950, 1957; BS 5930:1981). They can be summarized as providing data for the following. 1. Site selection. The construction of certain major projects, such as earth dams, is dependent on the availability of a suitable site. Clearly, if the plan is to build on the cheapest, most readily available land, geotechnical problems due to the high permeability of the sub-soil, or to slope instability may make the final cost of the construction prohibitive. Since the safety of lives and property are at stake, it is important to consider the geotechnical merits or demerits of various sites before the site is chosen for a project of such magnitude. 2. Foundation and earthworks design. Generally, factors such as the availability of land at the right price, in a good location from the point of view of the eventual user, and with the

3

Planning and Procurement

3.

4.

5.

6.

7.

planning consent for its proposed use are of over-riding importance. For medium-sized engineering works, such as motorways and multistorey structures, the geotechnical problems must be solved once the site is available, in order to allow a safe and economical design to be prepared. Temporary works design. The actual process of construction may often impose greater stress on the ground than the final structure. While excavating for foundations, steep side slopes may be used, and the in-flow of groundwater may cause severe problems and even collapse. These temporary difficulties, which may in extreme circumstances prevent the completion of a construction project, will not usually affect the design of the finished works. They must, however, be the object of serious investigation. The effects of the proposed project on its environment. The construction of an excavation may cause structural distress to neighbouring structures for a variety of reasons such as loss of ground, and lowering of the groundwater table. This will result in prompt legal action. On a wider scale, the extraction of water from the ground for drinking may cause pollution of the aquifer in coastal regions due to saline intrusion, and the construction of a major earth dam and lake may not only destroy agricultural land and game, but may introduce new diseases into large populations. These effects must be the subject of investigation. Investigation of existing construction. The observation and recording of the conditions leading to failure of soils or structures are of primary importance to the advance of soil mechanics, but the investigation of existing works can also be particularly valuable for obtaining data for use in proposed works on similar soil conditions. The rate of settlement, the necessity for special types of structural solution, and the bulk strength of the sub-soil may all be obtained with more certainty from back-analysis of the records of existing works than from smallscale laboratory tests. The design of remedial works. If structures are seen to have failed, or to be about to fail, then remedial measures must be designed. Site investigation methods must be used to obtain parameters for design. Safety checks. Major civil engineering works, such as earth dams, have been constructed over a sufficiently long period for the precise construction method and the present stability of early examples to be in doubt. Site investigations are used to provide data to allow their continued use.

According to US 5930: 1981, site investigation aims to determine all the information relevant to site usage, including meteorological, hydrological and environmental information. Ground investigation aims only to determine the ground and groundwater conditions at and around the site; this is normally achieved by boring and drilling exploratory holes, and carrying out soil and rock testing. In common engineering parlance, however, the terms site investigation and ground investigation are used interchangeably. GENERAL DESIGN PHILOSOPHY

Site investigation should be an integral part of the construction process. Unfortunately it is often seen as a necessary evil — a process which must be gone through by a designer if he or she is to avoid being thought incompetent, but one which gives little of value and takes precious time and money. This is an unfortunate by-product of the way in which site investigation is often carried out, and it can hardly be surprising that if no effort is put into targeting the investigation to precise issues, then little of value emerges. Site investigation should be a carefully considered process of scientific discovery, tailored both to the conditions existing on site and to the form of construction which is expected to take place. In order to make the most of site investigation, it is important that the design team (who may be led by architects, quantity surveyors and other non- engineering professionals) obtain at the conceptual design stage the advice, however briefly, of a geotechnical engineer. This geotechnical specialist can give the initial and most important guidance on the likely risks associated with the project, and the way in which they

4

Site Investigation may be investigated and dealt with. For most construction projects, the natural variability of the ground and groundwater conditions represent a major risk, which if not properly addressed can endanger not only the financial viability, but also the physical stability of the construction, either during construction or during the use of the building. In principal then, all sites must be investigated if construction is to be safe and economical. In practice, the way in which they are investigated can vary very widely, and the costs and time necessary will also be significantly different. The keys in selecting the most effective method of dealing with the inevitable uncertainties which must arise are geotechnical knowledge and experience. Possible approaches which have been successfully used include the following.

Approaches to site investigation Approach 1: Desk-study and geotechnical advice The minimum requirement for a satisfactory investigation is that a desk study and walk-over survey are carried out by a competent geotechnical specialist, who has been carefully briefed by the lead technical construction professional (architect, engineer or quantity surveyor) as to the forms and locations of construction anticipated at the site. This approach will be satisfactory where routine construction is being carried out in well-known and relatively uniform ground conditions. The desk study and walk-over survey (see Chapter 3) are intended to: 1. confirm the presence of the anticipated ground conditions, as a result of the examination of geological maps and previous ground investigation records; 2. establish that the variability of the sub-soil is likely to be small; 3. identify potential construction problems; 4. establish the geotechnical limit states (for example, slope instability, excessive foundation settlement) which must be designed for; and above all, to 5. investigate the likelihood of unexpected’ hazards (for example, made ground, or contaminated land). It is unlikely that detailed geotechnical design parameters will be required, since the performance of the proposed development can be judged on the basis of previous construction. Approach 2: ‘Standard’ ground investigation For most projects a more elaborate approach is needed, and will generally follow the following course. 1. A desk study and walk-over survey must first be carried out, to establish the likely conditions on and below the site, as described above (and see Chapter 3). 2. The details of the proposed construction must be ascertained, in as much as they have been decided. Particular care should be taken to establish the probable loading conditions and the sensitivity of any structures to be built, or those already existing on, around or below the site, to the changes that will occur as a result of construction. For example, services and tunnels passing below or alongside a proposed excavation for a basement may be damaged by the movements caused by excavation, and buildings above a proposed tunnel may be damaged by changes in the groundwater conditions and any ground loss caused by construction. 3. From the combinations of construction and ground conditions, the need for particular foundation types, for retaining walls, for cut slopes, and for special construction processes (such a grouting, dewatering and ground improvement) should be determined. These will bring with them particular limit states, and where limit states cannot be avoided (for example,

5

Planning and Procurement by changing the configuration of the proposed construction) there will be a need to carry out geotechnical analyses.

EXAMPLE: POTENTIAL LIMIT STATES • • • • • • • •

Bearing capacity failure of foundations Differential settlement of foundations leading to structural damage Instability of clay slopes Sulphate attack on concrete Damage by mining subsidence Damage to surrounding structures as a result of excavating or dewatering an excavation Ground collapse over pre-existing natural solution features Collapse of excavations as a result of excessive water inflow.

The identification of potential limit states is a matter of experience, education and pessimism. ‘Confidence may impress the Client, but it has little effect on the forces of nature’ (Skempton 1948). 4. At this stage the geotechnical designer for the project will need to estimate (from experience, or from published values, in papers, or from previous investigations in the same strata) the likely values of parameters required for analyses of limit states, for the various types of ground expected to occur at the site. Some preliminary geotechnical design of the project is required, in order to recognize that only a few of the possible limit states are likely to have to be faced, and therefore that more detailed investigations will not be required for many parameters: • where possible, limit states should be avoided, by choosing an appropriate form of structure (for example, by piling through soft clays, rather than designing for bearing capacity failure of shallow foundations); • it will be recognized that certain limit states will not be a problem (e.g. the bearing capacity of shallow foundations on rock). At this stage critical parameters, essential to the successful completion of the project, must be recognized. EXAMPLE: PARAMETERS REQUIRED FOR THE DESIGN OF A FOUNDATION IN CLAY • • • • • • •

Bulk unit weight of clay Undrained strength of clay Compressibility of clay Variability of the above, both laterally and with depth Groundwater level Sulphate content of groundwater Acidity of groundwater.

5. From a knowledge of the probable ground conditions and the required parameters, the geotechnical specialist should now identify all possible ways of determining the required parameters. Many tests that might be used (see Chapters 8 and 9) will only work satisfactorily in limited ground conditions, so limiting the available choice. In principle, the parameters may be obtained: • • •

based on published data from other sites; based on previous site investigation data; back-analysis of performance of nearby construction; 6

Site Investigation • • •

back-analysis of observed performance during construction; laboratory testing on samples taken during ground investigation; and in situ testing during ground investigation.

In order to optimize the investigation, estimates of: • • • •

relative accuracy; relative cost; availability, and relevance to the problem

should be assessed for each way of determining the parameters. EXAMPLE: DETERMINATION OF THE COMPRESSIBILITY OF FRACTURED WEAK ROCK SPT

Cheap, readily available, widely accepted, usable at any depth, inaccurate

Plate test

Expensive, readily available, accurate, widely accepted, difficult to use at depth

Surface-wave geophysics

Cheap, not readily available, relatively accurate, shallow only, not widely accepted

Back analysis

Virtually free, readily available, relatively accurate, any depth, may not be relevant if site conditions are unusual.

At the same time the degree of sophistication and the accuracy required for each type of geotechnical analysis should be determined. For ultimate limit states (i.e. where collapse is involved) consider the cost of failure, in terms of: • • •

legal; political; and financial consequences.

For serviceability limit states (i.e. where collapse does not occur, but the use of the structure is impaired) consider: • •

savings which might be made in construction costs if parameters were better known; and the reduction in risk that might be achieved by using better analytical methods, based upon sounder engineering, with more sophisticated parameters.

EXAMPLE: ESTIMATION OF GROUND MOVEMENTS AROUND DEEP EXCAVATIONS IN THE CITY OF LONDON Despite the generally large cost of civil engineering construction it is common to base routine design on basic parameters obtained from the SPT (Chapter 9), and routine undrained triaxial and oedometer testing (Chapter 8), both of which generally give very conservative (i.e. over-safe) estimates of ground movements. It is relatively unusual to base design on back-analysed parameters, or on more sophisticated and well-instrumented laboratory stress-path testing, despite the proven ability of these forms of parameters, in conjunction with finite element analysis, to give good predictions of movements around large excavations in the London clay. The cost of using higher quality ground investigation and analytical techniques is typically less than

7

Planning and Procurement 0.1% of the total cost of land purchase, architectural and structural design, and construction. Therefore it is worth considering whether these techniques may be used to justify greater site usage, such as building more basements and/or building closer to neighbouring structures. As will be seen below, increasing expenditure on geotechnical engineering can also be used to reduce the complexity of the construction process, which will lead directly to reductions in construction cost. 6. The details of the ground investigation can now be decided. The investigation boreholes should be of sufficient depth and distribution to establish the position of interfaces between different types of soil (within the zone likely to affect the construction), and the in situ testing and soil sampling should be planned so that the soil can be grouped into different categories (for example, rock, clay, sand, organic material — see Chapter 2) as well as tested to provide the specific parameters necessary for design calculations. This facet of the planning and design of a ground investigation is considered later in this chapter. Approach 3: Limited investigation, coupled with monitoring In some projects, it may be possible to carry out redesign during construction, in order to reduce costs. Given the natural variability of the ground, geotechnical engineers routinely use ‘moderately conservative’ soil parameters in design calculations, and do not normally attempt genuine predictions of values such as settlement, ground movements adjacent to excavations, etc. The example below illustrates how the demolition and reconstruction process was modified during construction, on the basis of moderately conservative design using finite element and boundary element analysis, and observations of ground movements.

EXAMPLE: SITE INVESTIGATION AND REDESIGN DURING CONSTRUCTION, FOR A BUILDING OVER A TUNNEL IN CENTRAL LONDON The development of Grand Buildings, in Trafalgar Square, London, required demolition and reconstruction techniques which could guarantee that damage to underlying Underground railway tunnels would be avoided (Clayton et al. 1991). The relative location of Grand Buildings, with respect to the underlying tunnels, can be seen in Fig. 1.1—the closest tunnels, approximately l0m in diameter, lie only 5m below the basement of the new building. It was thought that the effects of construction on the underlying tunnels would be acceptably small if ground movements at the tunnel level were less than 15 mm. Initial designs were based upon limited and rather routine’ ground investigation, involving just two boreholes. Strength and compressibility values were determined from standard triaxial and oedometer testing (see Chapter 8). These values were not, however, used in estimating ground movements around the structure, since it is known (Fig. 1.2) that in this part of the London clay deposit they very significantly underestimate the stiffness of the ground. Instead, the movements were calculated using finite element and boundary element computer methods, incorporating the ground stiffness values back analysed from observations of movements at the Hyde Park Cavalry Barracks, a site some distance away, but still in similar London clay. Even using these, much higher, stiffness parameters the estimated ground movements were large. In order to limit the predicted tunnel movements a complex 20-stage sequence of demolition and construction was developed, which involved construction of foundations from within the existing building, in a number of small areas, with underpinning, and the intermixing of construction with demolition, the provision of some kentledge to limit the effects of unloading, and extensive temporary works to support the partly demolished building.

8

Site Investigation

Fig. 1.1 Elevation showing a W—E section through Grand Buildings (above), and plan showing proposed demolition and raft construction sequence (below). During planning of the construction process it soon became apparent that the proposed sequence of demolition and reconstruction would prove very complex and time consuming in execution, and that therefore economies of time and cost might be achieved through a redesign. In the absence of goodquality site-specific soil parameters for use in further analyses, an observational approach was developed. This was not the Observational Method sensu stricto (see below), but a strategy based firmly upon measurement of a critical parameter, vertical displacement, at the level of the most critical tunnel. The strategy involved: 1. assessment of the available information on the London clay, including experience gained by the design-and-build contractor in constructing the adjacent Griffin House; 2. adoption of moderately conservative soil stiffness parameters, and a conservative demolition and reconstruction scheme starting at the least-sensitive (Griffin House) end of the existing Grand Buildings; 3. boundary element analysis to predict the movements at various levels beneath the structure, and especially at the most critical tunnel location, and along the Passenger Access Tunnel which runs at the same level from the Upper Machine Room towards Griffin House; 4. incorporation in the plans of elements of work which could be abandoned if the predicted ground movements were proved to be pessimistic;

9

Planning and Procurement 5. monitoring of movements within the Passenger Access Tunnel, especially during the early stages of demolition; and 6. re-assessment of ground stiffness parameters, and re-design of the demolition and reconstruction programme, as the demolition proceeded.

Fig. 1.2 Comparison of Young’s modulus values for the London clay at Grand Buildings, obtained from routine undrained triaxial and oedometer testing, with values back analysed from observed movements around other excavations in the London area. The resulting demolition areas (in numbered circles, according to sequence) are shown in Fig. 1.1. As a result of early measurements, during the demolition and excavation of strip 1, it became clear that the design analysis had significantly overestimated the heave. Therefore the planned ‘back-load’ kentledge was not used, except on strip 5 and immediately above the Upper Machine Room, and demolition was allowed to proceed simultaneously over the entire site. Monitoring continued throughout demolition and reconstruction. A maximum heave of the order of 4.3mm was measured, compared with values of the order of 10—15mm predicted by finite element and’ boundary element analyses for the original design. Approach 4: The observational method This is a carefully considered approach to geotechnical design, developed by Peck (1969).

Peck (1969) ascribed Terzaghi’s great success to his use of observation, coupled with his insistence on full, personal responsibility and authority on critical jobs. Clearly variations in financial constraints, the complexity of soil conditions, and time restrictions mean that very different approaches can be taken during site investigation. Peck argued that the methods available for coping with the inevitable uncertainties which arise as a result of the natural variability of soil and rock conditions broadly form three groups. 1. Method 1: Carry out limited investigation, and adopt an excessive factor of safety during design. 2. Method 2: Carry out limited investigation, and make design assumptions in accordance with general average experience.

10

Site Investigation 3. Method 3: Carry out very detailed investigation. In the first two methods only the vaguest approximations to the values of the physical properties of the sub-soil can be obtained. The variability of the soil properties, together with the degree of continuity of the individual layers of soil are almost certainly unknown, and groundwater conditions will not usually be adequately defined. Under these conditions it is almost certain that method 1 will be wasteful, while method 2 can frequently be dangerous. Only in the cases of investigations of major projects is there any likelihood that sufficient funds will be available for very detailed investigations, and in many cases the financial return will not merit this approach. Peck (1969) gives the ingredients of the ‘observational method’ as follows: 1. exploration sufficient to establish at least the general nature, pattern and properties of the deposits, but not necessarily in detail; 2. assessment of the most probable conditions and the most unfavourable conceivable deviations from these conditions. In this assessment geology often plays a major role; 3. establishment of the design based on a working hypothesis of behaviour anticipated under the most probable conditions; 4. selection of quantities to be observed as construction proceeds, and calculation of their anticipated values on the basis of the working hypothesis; 5. calculation of values of the same quantities under the most unfavourable conditions compatible with the available data concerning the subsurface conditions; 6. selection in advance of a course of action or modification of design for every foreseeable significant deviation of the observational findings from those predicted on the basis of the working hypothesis; 7. measurement of quantities to be observed and evaluation of actual conditions during construction; and 8. modification of the design to suit actual conditions. A simple example of the observational method is given by Peck (1969). The pressures applied by soil to a strutted excavation are, to this day, a matter of considerable uncertainty. Conventional design methods assume worst conditions, as determined by various instrumented sections (for example, Peck (1943)). The Harris Trust building was to be constructed in Chicago, and the contractor had to design a bracing system (Fig. 1.3) for the excavation for foundations. He had at his disposal various measurements of strut loads on similar ground in Chicago and could therefore predict with some certainty the maximum strut loads that would occur.

Fig. 1.3 The Harris Trust Excavation (Peck 1969).

11

Planning and Procurement

The design of the struts could have been based on the trapezoidal diagram, providing a safe but uneconomical design since most of the struts would have carried much less load than their capability. The contractor proposed to design the struts at a relatively low factor of safety, for loads of about twothirds the envelope values, or about average measured load conditions. This achieved considerable economy. To guard against higher loads the contractor measured the axial load in every strut during construction, and had available extra struts for immediate insertion if necessary. Only three struts were required in addition to the thirty-nine originally designed for the whole project. Not only did this approach produce a large saving in construction costs; it also, and perhaps more importantly gave the absolute certainty that no strut in the system was overloaded. The observational method is now frequently claimed to be used, when in fact all the essential components described above have not been adhered to. In 1985 Peck noted that: the observational method, surely one of the most powerful weapons in our arsenal, is becoming discredited by misuse. Too often it is invoked by name but not by deed. Simply adopting a course of action and observing the consequences is not the observational method as it should be understood in applied soil mechanics. Among the essential but often overlooked elements are to make the most thorough subsurface explorations that are practicable, to establish the course of action on the basis of the most probable set of circumstances and to formulate, in advance, the actions that are to be taken if less favorable or even the most unfavorable conditions are actually encountered. These elements are often difficult to achieve, but the omission of any one of them reduces the observational method to an excuse for shoddy exploration or design, to dependence on good luck instead of good design. Unhappily, there are far too many instances in which poor design is disguised as the state of the art merely by characterizing it as an application of the observational method.

IMPLEMENTATION

It has already been noted that early site investigation in Britain was associated with work by the Building Research Station, and by contractors. During this period the response of contractors rather than consultants in setting up geotechnical organizations meant that by the late 1940s a high proportion of the experience, expertise and facilities available for site investigation was held by contracting firms. As a result, site investigation in the UK became a contractual operation. At the present time, much of the work of site investigation is carried out on the basis of a competitive tender. At the present time then, most site investigation in Britain is commissioned by local authorities, government organizations or consulting engineers, on behalf of their clients. Typically the engineer produces conditions of contract, a specification, and a bill of quantities, and the tenderer receives a plan showing the proposed borehole locations. Provisional borehole depths and sampling routines are normally given, and the contractor will be told whether he is to provide a factual report, or whether both factual and interpretative reports on the project are required. Whether interpretation is required or not, it can be seen that the contractor is under great pressure to work quickly and efficiently, for the company will have quoted fixed prices for work to be carried out in uncertain ground and groundwater conditions. It has been found that the best site investigations involve a considerable number of activities, some of which may become relatively unimportant in some cases, but should never be forgotten. An ideal order of events might be as shown in Table 1.1.

The sequence of geotechnical site investigation might be: 1. preliminary desk study, or fact-finding survey;

12

Site Investigation 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.

air photograph interpretation; site walk-over survey; preliminary subsurface exploration; soil classification by description and simple testing; detailed subsurface exploration and field testing; the physical survey (laboratory testing); evaluation of data; geotechnical design; field trials; and liaison by geotechnical engineer with site staff during project construction.

Table 1.1 Order of events for site investigations Geotechnical designers Geotechnical contractor Definition of Geotechnical advice on likely design project issues Site selection Preliminary desk study to advise on relative geotechnical merits of different sites Conceptual design Geotechnical advice on optimizing structural forms and construction methods, in order to reduce sensitivity of proposed construction to ground conditions Detailed desk study and walk-over survey to produce a report giving: • expected ground conditions • recommended types of foundations • geotechnical design problems needing analysis

Project design team

Ground investigation plan Ground investigation: • profiling • classification • determination of parameters Detailed structural / architectural design

Detailed geotechnical design

Construction

Comparison of actual and anticipated ground conditions-assessment of new risks Geotechnical monitoring

Performance

Additional ground investigation Instrumentation

Unfortunately, in practice, British site investigation today more closely resembles the dream of engineers working on soil mechanics before World War I. According to Terzaghi (1936): engineers imagined that the future science of foundations would consist in carrying out the following program: Drill a hole into the ground. Send the soil samples obtained from the hole through a laboratory with standardized apparatus served by conscientious human automatons. Collect the figures, introduce them into equations, and compute the result.

13

Planning and Procurement

After a period of optimism between the wars, the inevitable pressures of competitive tendering have reduced the average level of British site investigation to the state where reputable companies, with considerable geotechnical experience and expertise to offer, find financial survival difficult. The financial pressures faced by British site investigation contractors are inevitable, whilst clients do not understand the value of good ground investigation, and prefer economy to sound engineering. The major part of the college training of civil engineers consists in the absorption of the laws and rules which apply to relatively simple and well-defined materials, such as steel or concrete. This type of education breeds the illusion that everything connected with engineering should and can be computed on the basis of a priori assumptions (Terzaghi 1936). PLANNING GROUND INVESTIGATIONS

The process of all site investigation should be, above all, one of scientific method. Sufficient factual information should be gathered (from the desk study and walk-over survey) to form hypotheses regarding the ground conditions, and from this and a reasonable knowledge of what is to be built on the site, the problems likely to be encountered both during the construction and the life of the development must be predicted. The design of the proposed construction should then, ideally, take into account the project’s geotechnical setting, in order to avoid as many difficulties as possible, and minimize the remainder. Finally, ground investigation should be carried out in order, if necessary, to determine the actual ground conditions on the site, and where necessary to obtain parameters for engineering calculations. Field investigation, whether by geophysics, or by boring or drilling, must have clearly identified aims if it is to be worthwhile. In some situations it may be necessary to make extensive and detailed ground investigations, but it is also perfectly conceivable that in other situations very few (if any) trial pits or boreholes or soil testing will be required before the start of construction. At present, ground investigation is poorly targeted, and it is because of this that it is sometimes regarded as a necessary but rather unrewarding expense. Yet it must be remembered that the majority of unforseen costs associated with construction are geotechnical in nature. Tyrell et al. (1983) carried out an appraisal of 10 UK highway construction projects where cost over-runs were substantial, averaging some 35% of the tender sum. They went through contract records to determine the cause of the additional costs, and found that approximately one-half of the increase in cost could be attributed to just two factors: 1. inadequate planning of ground investigation; and 2. inadequate interpretation of the results of ground investigations. Because the planning of ground investigation is so important, it is essential that an experienced geotechnical specialist is consulted by the promoter of the project and his leading technical designer very early during conceptual design (see Procurement, below). The planning of a ground investigation is broken down into its component parts in Table 1.2. The geotechnical specialist may be an independent consultant, but more often in the UK will work for a specialist geotechnical consultancy practice, for a - general civil engineering consultancy, or for one of the larger specialist ground engineering contractors. In the UK, the British Geotechnical Society’s 1992 Geotechnical Directory of the United Kingdom obtainable from the BGS at the Institution of Civil Engineers in London, gives a list of suitable individuals and the companies that employ them. The qualifications and experience required, before an individual may achieve an entry in the Directory, are shown in Table 1.3.

14

Site Investigation

Stage I II III IV V

VI VII VIII

Table 1.2 Planning a ground investigation Action Carried out by Obtain the services of an experienced geotechnical Developer/client specialist Carry out desk study and air photograph interpretation, Geotechnical specialist to determine the probable ground conditions at the site Conceptual design: optimize construction to minimize Architect, structural engineer, geotechnical risk geotechnical specialist Identify parameters required for detailed geotechnical Geotechnical specialist calculations Plan ground investigation to determine ground Geotechnical specialist conditions, and their variation, and to obtain geotechnical parameters Define methods of investigation and testing to be used Geotechnical specialist Determine minimum acceptable standards for ground Geotechnical specialist investigation work Identify suitable methods of procurement professional Geotechnical specialist, lead design, developer/client

The most important step in the entire process of site investigation is the appointment, at a very early stage in the planning of a construction project, of a geotechnical specialist. At present, much site investigation drilling and testing is carried out in a routine way, and in the absence of any significant plan. This can result in a significant waste of money, and time, since the work is carried out without reference to the special needs of the project. Table 1.3 Requirements for organizations and individuals to appear in the British Geotechnical Society’s Geotechnical Directory of the UK Organizations

Individuals

For an organization to appear in the Directory it must be active in the UK offering services in geotechnical engineering (as opposed to manufacturing geotechnical equipment, for example). It must also employ at least one person whose name appears as an individual entry in the Directory.

For an individual’s name to appear in the Directory, he or she must be resident in the UK and be a member of the British Geotechnical Society, the Engineering Group of the Geological Society or a regional Geotechnical Society.

Organizations which belong to one of the Trade Associations featured in the Directory are identified in the lists by means of the Association’s logo.

He or she must also fulfil one of the three sets of criteria given below. Chartered Engineer through Corporate Membership of the Institution of Civil Engineers, the Institution of Structural Engineers, the Institution of Mining and Metallurgy or be a Chartered Geologist or a Corporate member of an equivalent overseas Institution and a minimum of five years’ experience as a practising geotechnical specialist or: a professional qualification, as above and a further degree (a Master’s degree or Doctorate) in a relevant subject area, for example soil mechanics, geotechnical engineering, foundation engineering or engineering geology and a minimum of three years’ experience as a practising geotechnical specialist or: a minimum of twenty years’ experience as a practising geotechnical specialist

Once a geotechnical specialist has been appointed, work can start on determining the ground conditions at the site. The first stages of this process are the desk study, air photograph interpretation, and a site walk-over survey (see Chapter 3). In geotechnical work, descriptions of soil and rock are made in accordance with very specific guidelines (Chapter 2), which have been devised to indicate their performance under engineering conditions, in terms of strength, compressibility and

15

Planning and Procurement permeability. If previous site investigation reports exist for construction in the same soil, this allows the geotechnical engineer to judge (albeit in a general way) the likely performance of the ground under and around the proposed development. In any case, geological maps coupled with experience will give a considerable amount of information, of great value in the initial stages of design. At this stage there should also be interaction between the client and all of his design professionals. Where possible, the design should be modified to reduce possible geotechnical problems. For example, if a large site is to be developed as a business park, the buildings might be re-aligned with their long sides parallel to the contours; this will reduce the amount of cut and fill, thus keeping the cost of foundations and retaining structures to a minimum, while also reducing the risks of slope instability. Structures may be relocated to avoid areas of potentially difficult ground, such as infilled quarries, pre-existing slope instability, or where old foundations or contaminated ground may exist below previously demolished structures. Appropriate foundation types and structural connexions can be chosen. From a knowledge of the probable ground and groundwater conditions, and the required structural form(s), the geotechnical engineer should predict the types of foundations and earth-retaining structures required on the project, and any possible problems (such as slope instability, chemical attack on foundation concrete, construction difficulties) which can be foreseen, and which may therefore require further investigation. The planning of a ground investigation requires a knowledge both of the ground conditions at and around the site, and of the form of the proposed construction. If the design of the construction is to be optimized, then the form of construction should, as far as possible, take the expected ground conditions into account. At the end of the desk study, air photo interpretation and walk-over survey, the geotechnical specialist should make a written report, giving he expected ground conditions across and around the site, the uncertainties in these predictions, and the extent of ground investigation proposed for their investigation. In addition, he or she should make proposals for suitable types of foundations for any proposed structures, and should identify areas where other geotechnical structures (such as retaining walls or slopes) will be expected. For these areas there will be a need to obtain geotechnical parameters for design. Other potential problems requiring investigation should also be identified. The parameters to be obtained during ground investigation, and the methods to be used to obtain those parameters, should be described, and justified, in detail. Planning trial pitting, boring and drilling Drilling and trial pitting are normally carried out for a number of reasons, such as: 1. 2. 3. 4. 5. 6.

to establish the general nature of the strata below a site; to establish the vertical or lateral variability of soil conditions; to verify the interpretation of geophysical surveys; to obtain samples for laboratory testing; to allow in situ tests to be carried out; and to install instruments such as piezometers, or extensometers.

Frequently, most if not all of these objectives will control the method of drilling on site. All the objectives must be achieved with the minimum of expense and disruption to occupiers of the site. In the UK, drilling, sampling and testing are normally carried out by a specialist site investigation contractor. The most convenient method of organizing the work is for the engineer controlling the contract to decide on the position and depth of boreholes, the sampling routine for each soil type that is likely to be found, and the number and type of in situ and laboratory tests that are required. A number of contractors can then provide competitive bids, the cheapest price can be selected, and the work carried out.

16

Site Investigation

The scheme described in the preceding paragraph is ideal from the contractual viewpoint, because it allows a fixed price to be obtained by competitive tender. As a method of achieving the aims of site investigation, it is rarely satisfactory, however, because soil conditions are not very well known at tender stage and because competitive tendering favours contractors who have the lowest overheads and are therefore less likely to be able to bring a high level of engineering expertise to bear on the work. When specifying and controlling drilling, it is important that the drilling and testing programme can be modified while work is in progress, as new information is made available by each borehole or test pit. Therefore both office and site staff should be aware of the reasons for the decisions made during the initial planning of the work, in order that they do not hesitate to alter drilling and testing schedules where this is appropriate. The principal factors which allow a logical drilling programme to be planned and successfully executed are: 1. a relationship between structure, borehole layout, frequency and depth; 2. a need for sample quality and quantity related to the required geotechnical parameters and the soil type and variability; 3. site supervision, to ensure that drilling and sampling are carried out to a high standard and that good records are kept; and 4. prompt sample description and preparation of borehole and pit records in order that the drilling programme can be modified as the work proceeds. These factors are considered in turn below.

Borehole layout and frequency Borehole layout and frequency are partly controlled by the complexity of the geological conditions. The complexity of geological structure and the variability of each of the soil or rock units should be at least partially known after the fact-finding or desk study. If soil conditions are relatively uniform, or the geological data are limited, the following paragraphs will give an initial guide. Borehole layout and frequency may need to be changed as more information emerges. Investigation will normally be carried out by machine or hand-excavated trial pits, where only shallow depths are to be investigated, for example for low-rise housing projects, or for shallow instability problems. The use of pits in these situations allows a detailed engineering description of soil conditions, and will also permit block samples to be taken. Most boreholes will be considerably deeper than can be excavated by an open trial pit, and these will normally be carried out by light percussion or hollow stem auger drilling.

Most projects will fall into one of the following categories: a) isolated small structures, such as pylons, radio masts, or small houses, where one borehole may be sufficient; b) compact projects, such as buildings, dams, bridges or small landslips, will require at least four boreholes. These will normally be deep and relatively closely spaced; c) extended projects, such as motorways, railways, reservoirs and land reclamation schemes will require shallower, more widely spaced boreholes, but these will normally be expected to verify the depth of ‘good’ ground. In the case of road projects this will mean either rockhead, or a soil with a ‘stiff’ consistency. In the case of reservoirs, borings should be continued until an adequate thickness of impermeable ground is found. The frequency of borings on extended sites must be judged on the basis of the uniformity or otherwise of the site geology and its expected soil variability. On a highway project the recommendations for borehole spacing

17

Planning and Procurement vary from 30 to 60 m (Hvorslev 1949) to 160 m in changeable soils and 300m in uniform soils (Road Research Laboratory 1954). Many projects, such as highways, are a combination of the categories described above. Structures on extended projects should be treated as compact projects. For example, a typical investigation for a motorway in the UK might use 5—l0m deep borings every 150m along the proposed road line, with four 25—30m deep borings at the proposed position of each bridge structure. Additional boreholes might be placed on the basis of soil information found during the fact-finding survey, on the basis of: 1. the geological succession in the area. Thin beds of limited outcrop may require closer boreholes; 2. the presence of drift deposits such as alluvium or glacial till, whose vertical and lateral extent may require close inspection; 3. problem areas, for example where pre-existing slope instability is suspected. The layout of the borings should aim not only to provide soil profiles and samples at positions related to the proposed structures and their foundations, but should also be arranged to allow the hypotheses formed during the fact-finding survey to be checked. The borings should be positioned to check the geological succession and to define the extent of the various materials on site, and they should be aligned, wherever possible, in order to allow cross-sections to be drawn (Fig. 1.4). Where structures are to be found on slopes, the overall stability of the structure and the slope must obviously be investigated, and to this end a deep borehole near the top of the slope can be very useful.

Fig. 1.4 Alignment of boreholes.

18

Site Investigation Depth of borings It is good practice on any site to sink at least one deep borehole to establish the solid geology. On extended projects several of these may be necessary, partly in order to establish the depth of weathering, which may be up to 100 m below ground level and may be irregular, and also to establish the depth to which cavernous or mined areas descend. Hvorslev (1949) suggested a number of general rules which remain applicable: The borings should be extended to strata of adequate bearing capacity and should penetrate all deposits which are unsuitable for foundation purposes — such as unconsolidated fill, peat, organic silt and very soft and compressible clay. The soft strata should be penetrated even when they are covered with a surface layer of high bearing capacity. When structures are to be founded on clay and other materials with adequate strength to support the structure but subject to consolidation by an increase in the load, the borings should penetrate the compressible strata or be extended to such a depth that the stress increase for still deeper strata is reduced to values so small that the corresponding consolidation of these strata will not materially influence the settlement of the proposed structure. Except in the case of very heavy loads or when seepage or other considerations are governing, the borings may be stopped when rock is encountered or after a short penetration into strata of exceptional bearing capacity and stiffness, provided it is known from explorations in the vicinity or the general stratigraphy of the area that these strata have adequate thickness or are underlain by still stronger formations. When these conditions are not fulfilled, some of the borings must be extended until it has been established that the strong strata have adequate thickness irrespective of the character of the underlying material. When the structure is to be founded on rock, it must be verified that bedrock and not boulders have been encountered, and it is advisable to extend one or more borings from 10 to 20 ft into solid rock in order to determine the extent and character of the weathered zone of the rock. In regions where rock or strata of exceptional bearing capacity are found at relatively shallow depths — say from 100 to 150 ft — it is advisable to extend at least one of the borings to such strata, even when other considerations may indicate that a smaller depth would be sufficient. The additional information thereby obtained is valuable insurance against unexpected developments and against overlooking foundation methods and types which may be more economical than those first considered. The depth requirements should be reconsidered, when results of the first borings are available, and it is often possible to reduce the depth of subsequent borings or to confine detailed and special explorations to particular strata.

As a rough guide to the necessary depths, as determined from considerations of stress distribution or seepage, the following depths may be used. 1. Reservoirs. Explore soil to: (i) the depth of the base of the impermeable stratum, or (ii) not less than 2 x maximum hydraulic head expected. 2. Foundations. Explore soil to the depth to which it will be significantly stressed. This is often taken as the depth at which the vertical total stress increase due to the foundation is equal to 10% of the stress applied at foundation level (Fig. 1.5). 3. For roads. Ground exploration need generally only proceed to 2—4 m below the finished road level, provided the vertical alignment is fixed. In practice some realignment often occurs in cuttings, and side drains may be dug up to 6 m deep. If site investigation is to allow flexibility in design, it is good practice to bore to at least 5 m below ground level where the finished road level is near existing ground level, 5 m below finished road level in cut, or at least one-and-ahalf times the embankment height in fill areas. 4. For dams. For earth structures, Hvorslev (1949) recommends a depth equal to one-half of the base width of the dam. For concrete structures the depth Of exploration should be between one-and-a-half and two times the height of the dam. Because the critical factor is safety against seepage and foundation failure, boreholes should penetrate not only soft or unstable materials, but also permeable materials to such a depth that seepage patterns can be predicted.

19

Planning and Procurement 5. For retaining walls. It has been suggested by Hvorslev that the preliminary depth of exploration should be three-quarters to one-and-a-half times the wall height below the bottom of the wall or its supporting piles. Because it is rare that more than one survey will be carried out for a small structure, it will generally be better to err on the safe side and bore to at least two times the probable wall height below the base of the wall. 6. For embankments. The depth of exploration should be at least equal to the height of the embankment and should ideally penetrate all soft soils if stability is to be investigated. If settlements are critical then soil may be significantly stressed to depths below the bottom of the embankment equal to the embankment width.

Fig. 1.5 Necessary borehole depths for foundations. Because many investigations are carried out to determine the type of foundations that must be used, all borings should be carried to a suitable bearing strata, and a reasonable proportion of the holes should be planned on the assumption that piling will have to be used.

Sampling, laboratory testing and in situ testing requirements As will be seen in the Chapters 6, 7 and 9, which deal with sampling disturbance, sampling techniques, and in situ testing, most available sampling and in situ testing techniques are imperfect, and often represent a compromise. The normal sampling and in situ testing routines in use in the UK, represent the results of just such a compromise. They result from the fact that stiff clays, stoney glacial

20

Site Investigation tills and gravelly alluvium are so often found in the UK, and that prices for ground investigation are relatively low. In routine ground investigations samples are taken or in situ tests made only every 1.5 m down boreholes, and only about 25% of the soil at every borehole location is sampled, however imperfectly. Even in the most intensely investigated site, it is unlikely that more than one part in 1000000 of the volume of ground affected by construction will be sampled. The sampling routine should be aimed at: 1. providing sufficient samples to classify the soil into broad soil groups, on the basis of particle size and compressibility; 2. assessing the variability of the soil; 3. providing soil specimens of suitable quality for strength and compressibility testing; and 4. providing specimens of soil and groundwater for chemical testing. Soil and rock are not normally found in pockets, each of a distinct type, but often grades gradually from one soil type (for example, sand) to another (for example, clay). It is therefore necessary artificially to divide the available soil and rock samples into groups, each of which is expected to have similar engineering behaviour. Engineering soil and rock description (Chapter 2), and index tests and classification tests are used for this purpose (Chapter 8). Geotechnical parameters are obtained by testing specimens which have been selected to be representative of each of the soil groups defined by soil description, and classification and index testing. Where soil grouping cannot be carried out, perhaps because of time or financial constraints, it is often found to be necessary to carry out much larger numbers of the more time-consuming and sophisticated tests required for determining geotechnical design parameters. Therefore this is a false economy. Thus, if 450 mm long samples are to be taken every 1.0 to 1.5 m down the borehole in cohesive soils, every test specimen should be subjected to determinations of water content and plasticity. Where an undrained shear strength profile is required, tests should be made on every specimen of the appropriate diameter in the depth range required for the profile. For proposed spread foundations, embarkments and temporary works cuttings, these depths should not be less than the height of the cut or fill, or the width of the foundation. If soil conditions are unfavourable, piles may be required; in anticipation of this, shear strengths should then be determined to much greater depths. Large numbers of undrained triaxial strength tests are required in order to establish a shear strength — depth profile in firm to hard clays, because of the scatter in their results which is induced by fissuring. In the past, it has often been assumed that much smaller numbers of effective strength test results will be needed, because fissuring effects are less important. It now appears that this is not the case. Fissures appear to have little effect on small-strain stiffness, but unfortunately give rise to a large scatter in effective strength parameters (c’ and p’) even when 100mm diameter specimens are used. Current UK practice tends to underestimate the need for a sufficient number of effective stress tests; when longterm slope or retaining wall stability problems must be analysed, at least five sets of tests, each with three specimens, should be made on each soil type. Compressibility tests, normally by oedometer compression, will be required from every specimen within the probable depths of soil to be significantly stressed. Clearly, soil is normally variable, and when a two-stage investigation (a variation survey followed by detailed exploration) is not carried out, the only logical course is to test more extensively those specimens that are obtained. In the UK in situ testing is carried out when: 1. good quality sampling is impossible (for example, in granular soils, in fractured rock masses, in very soft or sensitive clays, or in stoney soils); 2. the parameter required cannot be obtained from laboratory tests (for example, in situ horizontal stress);

21

Planning and Procurement 3. when in situ tests are cheap and quick, relative to the process of sampling and laboratory testing (for example, the use of the SPT in London clay, to determine undrained shear strength); and most importantly, 4. for profiling and classification of soils (for example, with the cone test, or with dynamic penetration tests). The most commonly used test is the Standard Penetration Test (SF1’) (Chapter 9), which is routinely used at 1.5 m intervals within boreholes in granular soils, stoney soils, and weak rock. Other common in situ tests include the field vane (used only in soft and very soft cohesive soils), the plate test (used in granular soils and fractured weak rocks), and permeability tests (used in most ground, to determine the coefficient of permeability). Marsland (1986) has stated that: the choice of test methods and procedures is one of the most important decisions to be made during the planning and progress of a site investigation. Even the most carefully executed tests are of little value if they are not appropriate. In assessing the suitability of a particular test it is necessary to balance the design requirements, the combined accuracy of the test and associated correlations, and possible differences between test and full-scale behaviour.

The primary decision will be whether to test in the laboratory or in situ. Table 1.4 gives the relative merits of these options. Table 1.4 Relative merits of in situ and laboratory testing In situ testing Laboratory testing Advantages Test results can be obtained during the course of Tests are carried out in a well-regulated environment. the investigation, much earlier than laboratory test results Stress and strain levels are controlled, as are drainage boundaries and strain rates. Appropriate methods may be able to test large volumes of ground, ensuring that the effects of Effective strength testing is straightforward. large particle sizes and discontinuities are fully represented. The effect of stress path and history can be examined. Estimates of in situ horizontal stress can be obtained. Drained bulk modulus can be determined. Disadvantages Drainage boundaries are not controlled, so that it Testing cannot be used whenever samples of cannot definitely be known whether loading tests sufficient quality and size are unobtainable, for example, in granular soils, fractured weak rock, are fully undrained. stoney clays. Stress paths and/or strain levels are often poorly Test results are only available some time after the controlled. completion of fieldwork. Tests to determine effective stress strength parameters cannot be made, because of the expense and inconvenience of a long test period. Pore pressures cannot be measured in the tested volume, so that effective stresses are unknown. The ground investigation planner requires a detailed and up-to-date knowledge of both laboratory and in situ testing, if the best choices are to be made. Table 1.5 gives a summary of the current situation in

22

Site Investigation the UK — but this will rapidly become out of date. Whatever is used depends upon the soil and rock encountered, upon the need (profiling, classification, parameter determination), and upon the sophistication of geotechnical design that is anticipated.

Purpose Profiling

Classification

Table 1.5 Common uses of in situ and laboratory tests Suitable laboratory test Suitable in situ test Cone test Moisture content Dynamic penetration test Particle size distribution Plasticity (Atterberg limits) Geophysical down-hole logging Undrained strength Particle size distribution Cone Plasticity (Atterberg limits)

Parameter determination: Undrained strength, cu Peak effective strength, c’ φ’ Residual strength, cr’ φr’ Compressibility

Permeability Chemical characteristics

Undrained triaxial

SPT Cone Vane

Effective strength triaxial Shear box Ring shear Oedometer Triaxial, with small strain measurement Triaxial consolidation Triaxial permeability pH Sulphate content

Self-boring pressuremeter Plate test In situ permeability tests Geophysical resistivity

Geophysics Geophysical methods (Chapter 4) may be used for: 1. geological investigation, for example in determining the thickness of soft, superficial deposits, and the depth to rock, and in establishing weathering profiles, usually to provide crosssections; 2. resource assessment, for example the location of aquifers, the delineation of saline intrusion, the exploration of the extent of sand and gravel deposits, and rock for aggregate; 3. detecting critical buried features, such as voids (mineshafts, natural cavities, adits, pipelines) and buried artefacts (old foundations, wrecks at sea, etc.); and 4. determining engineering parameters, such as dynamic elastic moduli, and soil corrosivity. In some instances (for example, the determination of small-strain stiffness) they may be used in the same way as other in situ tests, but generally they are used as a supplement to direct methods of investigation, carried out by boreholes and trial pitting. The planning of geophysical surveys has been described in detail by Darrocott and McCann (1986). They note that clients have often voiced their disappointment with the results of geophysical site investigation, and note that in their experience the failure of the techniques can usually be attributed to one or more of the following problems: 1. inadequate or bad planning of the survey; 2. incorrect choice or specification of the technique;

23

Planning and Procurement 3. the use of insufficiently experienced personnel to conduct the survey. Geophysical surveys should be planned as an integral part of the site investigation. The desk-study information must be available so that the most effective techniques are used, and (as with direct methods of investigation, such as boring and trial pitting) the ‘targets’ of each part of a geophysical survey must be clearly understood. Table 1.6 shows how a geophysical survey should be planned. This is discussed in more detail in Chapter 4. Table 1.6 Stages of a geophysical survey as part of a ground investigation Stage I

Action Preliminary meeting between geophysicist and geotechnical specialist

II

Carry out desk study

III

Plan geophysical survey

IV V VI

Carry out geophysical trials Main geophysical survey On-site interpretation

VII

Correlation boreholes

VIII

Final interpretation

IX

Reporting

X

Feedback

Details Determine: (a) the precise result(s) expected (the ‘targets’); (b) whether geophysical methods can be expected to achieve (a); (c) which technique(s) are likely to be successful; (d) consider cost-effectiveness of geophysics relative to other techniques. Determine: (a) ground conditions; (b) groundwater conditions; (c) sources of background interference. Determine: (a) which techniques are likely to be successful, given the ground conditions, the targets’,and the background interference; (b) probability of success with each technique; (c) cost-effectiveness of geophysics relative to other techniques; (d) if geophysics appears possible, chose equipment and plan layout for chosen techniques, and identify suitable personnel. This will only be possible in unusual circumstances. The plan for the geophysical survey (for example, the layout of instruments) may need revision in the light of early data, to improve results The borehole programme should include holes to allow checking and ‘calibration’ of the geophysics. If possible these data should be made available to the geophysical survey team during their field work. Final interpretation should be made jointly by experienced geophysicists and geotechnical engineers, drawing together all the data, including that from direct investigation methods. Reporting should include raw data, in electronic form, as well as filtered, processed and interpreted results. The success of the work, as found during construction, should be conveyed to the geophysical team.

Specification As noted in Table 1.2, it is necessary to define, in one way or another, the minimum standards of the work to be carried out during the ground investigation. This is particularly important for all elements of work that are to be procured on the basis of competitive tender, since the specification document is central to the prices offered by contractors when bidding. The principal features of the specification contract documents in common use in the UK are given below.

24

Site Investigation 1. Entry, access and reinstatement. Whilst the engineer is responsible for arranging access, the contractor must give sufficient notice of entry. Only agreed access routes to the site of the boreholes can be used, and avoidable damage must be made good by the contractor at his own expense. The contractor must include in his rates for stripping topsoil in the area of the borehole, and for making good damage in the area of the borehole and along the access route. Unavoidable damage to crops and hedges or fences is normally paid for by the client. 2. Services. Services are to be located by hand digging a pit 1.5 m deep, where it is thought that service pipes, cables or ducts may be present in the area of a borehole. Precautions should be taken to protect field personnel from safety hazards, such as underground electrical cables and gas pipes. Engineers involved in ground investigation should recognize that they are responsible for the safety of those working for them. Public utility companies (gas, electricity, telephone, water, etc.) must be contacted to ensure that, as far as possible, risks to health and safety are properly identified before drilling is started. 3. Trial pits. The contractor should excavate trial pits by hand or machine in order that soil can be examined in situ and samples taken. The plan area of any such excavation should not normally be less than 2m2. Pits should be kept free from water, where encountered, by pumping. The contractor should supply, fix and remove, on completion, sufficient support to the side of the pits to protect anyone entering the hole. Topsoil should be stripped from the pit area before the start of work and should be stockpiled separately until completion. At the end of work, the pit should be filled with compacted spoil, any surplus being heaped proud over the site and covered with the topsoil. Where pits must be left open overnight the contractor must provide temporary fencing around the excavation. 4. Boring and drilling. For light percussion boring the minimum borehole diameter is normally 150 mm, but the contractor is responsible for starting the hole at a sufficiently large size to allow him to complete the hole to the required depth. If he fails to do this, the contractor is responsible for reboring the hole at his own expense. Claycutters should not be used in soft alluvial soils, where they may cause significant disturbance ahead of the hole. In some specifications the weight of the claycutter (see Chapter 6) has been limited as shown in Table 1.7. Shells used for boring in granular soils must not be tight fitting if this causes the soil to blow into the base of the hole. Under these conditions the borehole must be kept full of water at all times, and the shell should have a diameter not more than 90% of that of the inside of the borehole casing. Table 1.7 Specifications for claycutter Maximum weight of Diameter of claycutter and sinker boring (mm) bar (kg) 150 150 200 180 In a document produced by the Association of Ground Investigation Specialists in 1979, it was specified that a shell diameter at least 25mm less than that of the casing should be used (AGIS 1979), and in the current British Standard for the SPT it is a requirement that the outside diameter of the drilling tools should not exceed 90% of the inside diameter of the casing. The addition of water to borings is variously specified, with some documents preventing the addition of water except in ‘dry granular soils and stiff clays’. In one document the limit for addition of water to clays is fixed by testing the ‘immediate undrained cohesive strength’ with a ‘small field penetrometer’. Water can only be added to the borehole if the strength exceeds 140 kN/m2. Ideally, water should not be added to boreholes when drilling in clays above 25

Planning and Procurement groundwater level, whatever their consistency. Once the groundwater table is reached, then rapid drilling in stiff fatty clays may not allow time for swelling to take place. If this action is adopted, the first 1.0 m of the day’s drilling should not be sampled as it will have had time to swell as a result of stress relief. In all soils below the water table, the borehole should be kept full of water or drilling mud in order to reduce the effects of stress relief. In very soft or soft soils, this is also necessary to prevent failure of the soil up into the cased borehole. When casing is used, it should never be advanced ahead of the borehole. The bottom of the casing should preferably be kept within 150mm of the bottom of the hole at all times in order to prevent excess loosening of surrounding soil, or the formation of voids. In soft ground, both light percussion boring and auger boring are normally acceptable in the UK. Washboring or jetting is not permitted. Rotary core drilling may be carried out by open-holing through soft materials, or by drilling ahead of a soft ground boring which has already been made. Clauses are normally included to point out that the material to be cored may be friable or soft, or may contain mixtures of hard rock with interlayered soft materials. The contractor is normally responsible for selecting equipment which will satisfy the other requirements of the specification (for example, recovery, diameter, etc.). Some specifications require the use of hydraulic feed rigs, which are in almost exclusive use in the UK. The introduction of top drive rigs, however, has the advantage that larger runs can be made without rechucking. The Bill of Quantities is often arranged so that rotary core payment can be made for openholing and for the recovery of core. In this case it seems sensible to specify that the contractor shall use the necessary equipment, feed pressures and rates, and run lengths so that 100% recovery in any run can be achieved. Where less than 80% recovery is obtained, payment should be at the rate for open-holing for that run length. Drilling equipment should in general conform to BS 4019, although sampler barrels other than the double-tube ball-bearing swivel type with knock-on spring core catcher box and face discharge bit should not be discouraged since these may give good results. The minimum core diameter and the depth to which it is to be used should be specified, since the cost of deep larger diameter holes will increase significantly when small highly mobile rigs are in general use. The minimum size in sound rock should be N (or 76mm metric), with H (or 101 mm) used in soft or highly weathered rocks, and P (or 116 mm) used in drift such as stiff clays and glacial till. The maximum run length should be 3m, but where recovery is reduced to less than 80% the length of the next run is often specified as 1 m. If blocking of the flush ports or loss of flush return is detected at any stage, the barrel and core must be removed from the hole immediately. Clear water is the normally specified flush-fluid in the UK, with bentonite mud sometimes being specified in glacial drift and compressed air being used in soft rocks where water flush causes serious deterioration of the core. There has been a trend in the UK, in recent years, towards the use of rotary coring to obtain samples of heavily overconsolidated clays, and for this purpose bentonite or polymer muds are sometimes specified. All boreholes and drillholes should be backfilled and compacted in such a way that subsequent settlement of the backfill is avoided. Under artesian groundwater conditions, special sealing devices may be required. 5. Sampling. The contractor is commonly required to take disturbed samples, open-drive samples, piston-drive samples and rotary core from boreholes and drillholes. All samples from

26

Site Investigation soft ground borings or trial pit excavations should be clearly labelled with the following information: (i) contract name and reference number; (ii) reference number of hole; (iii) reference number of sample; (iv) date of sampling; (v) depth of top and bottom of sample below ground level; and (vi) top (if undisturbed). In addition to labelling the outside of the sample tube, a similar label, but additionally marked ‘top’ should be placed inside the top of the tube. All labelling should be protected from the effects of damp and water. Small disturbed samples are normally specified at the top of each stratum, from between undisturbed samples in fine-grained soils, and from the cutting shoes of all thick-walled opendrive samplers. They should contain not less than 1kg of soil which should, as far as possible, fill an airtight container. Large disturbed samples are normally taken from the test section of borehole used for the SPT (cone) test in gravels and other materials containing coarse particles. Their minimum weight should be 25 kg, although larger samples may be required for specific testing requirements. Thick-walled open-drive ‘undisturbed’ samples are standard in firm to very stiff clays in the UK. Most specifications make reference to the British Standard for Site Investigation, and in addition some specify minimum sample tube length (450mm), maximum area ratio (about 25%), inside diameter (100mm) and cutting edge taper ( >/ 20°). The cutting edge should be sharp and free from burrs. The sample tube and cutting shoe should be free from rust, pitting or burring. The use of oil on the inside of the tube should be limited to the minimum practicable. Thick-walled open-drive sample tubes should either be jacked into the ground or driven from ground level using a standard penetration test automatic trip hammer. Before lowering the sampler tube down the hole, the bottom of the boring should be cleaned of loose materials. Under extreme circumstances, the use of hand-rotated augering is specified for the 600mm of boring above the sample depth. In order to improve recovery, specifications sometimes require either sampler rotation (if practicable) or a waiting period in order to increase adhesion between the soil and the inside of the sampler tube. Thick-walled open-drive samplers must have a ball valve fitted in the sampler head to prevent the build up of pressure over the sample during the sample drive. This should be kept clean at all times. Flap-type core catchers inserted between the cutting shoe and sample tube are normally only permitted when all else fails. Over-driving should normally be avoided. The undisturbed sample should be pulled slowly from the soil and brought to the top of the hole. After removing the cutting shoe and the head, disturbed material from the top of the sample should be removed and sufficient soil taken from the base of the tube to allow a 10mm thick wax seal to be placed. The sample should immediately be sealed with at least three brush-coated layers of low melting point microcystalline wax. Following this, an oversize metal foil disc is sometimes specified, which is then covered with further wax. The ends of the tubes should be filled with a damp packing material (sawdust, newspaper or rags), and metal or plastic caps applied. The sample tube should be labelled immediately. In the UK, thickwalled open-drive samples are normally specified at a minimum of one every 1.0m for the first 5 m of drilling, and 1.5 m thereafter. Piston-drive sampling is normally only vaguely specified, usually for very soft or sensitive soils: ‘Equipment shall be of a pattern approved by the Engineer’. If piston samples are required then the equipment should be of the fixed piston type, and samples should be taken

27

Planning and Procurement continuously or at 1 m intervals. Clearly, much more stringent specifications are required for sampling sensitive soils, and therefore the following points should be included. Piston samples should be of the fixed piston type, with an area ratio compatible with their cutting edge angle (see the ISSMFE recommendations (1965) in Chapter 6). The maximum cutting edge angle should be 7°. In alluvial soils the minimum diameter should be 100mm, with a minimum length of 450 mm. The maximum inside clearance should be 0.75—1 .00%, although in very soft and sensitive soils there will be no necessity to include any inside clearance. Where possible, piston samplers should be of a design using short sectional liners made of an inert substance such as plastic or impregnated paper. They may be pushed to the desired sample depth or used from the base of a borehole. During the sample drive the inner (piston) rod must be securely fastened at ground surface so that no downward movement is possible. After sampling, the sampler should be rotated before being carefully brought to the surface. The liners should be removed, immediately labelled, and sealed with wax and pushon caps. When undisturbed sampling is attempted but no recovery results, the borehole should be cleaned out to the full depth to which the sampler has penetrated, and a fresh attempt to sample should be made immediately. The disturbed soil removed from the borehole should be saved as a large disturbed sample. In some specifications reduced payment is made to the contractor for undisturbed sampling attempts which give samples of less than 100mm length, or if the sample is of no use, provided the contractor is not at fault. When full recovery is not achieved the actual sample length and reason for partial recovery must be recorded. Rotary core should not be removed from the core-barrel by suspending it from a winch rope and hammering the inner barrel. Corebarrels should be held horizontally whilst cores are extruded using a coreplug by applying a constant pressure, and the cores should leave the barrel and travel on a transparent polythene sheet placed on a rigid plastic receiving channel of approximately the same diameter as the core. After extrusion the core should be sealed in the plastic sheet with waterproof adhesive tape, and bound to the rigid plastic receiving channel. It should then be placed in a corebox such as shown in Fig. 1.6. Wooden spacer blocks should indicate the top and bottom levels of each run.

Fig. 1.6 General layout for a corebox (from Redford 1981). Alternatively, a clear rigid plastic liner (in the UK, sometimes sold under the brand name ‘Coreline’) may be used as a third liner within the corebarrel. This reduces frictional forces between the inner barrel and the core, and in addition allows the core to be withdrawn from the inner barrel whilst it is held in the horizontal position. When a clear rigid plastic liner is used, end caps and plastic tape can be used to protect the core, and coreboxes need not be so carefully made.

28

Site Investigation

In general, it is common to see considerable detail relating to boring and drilling in specification documents, simply because there are currently no national or international standards available for guidance. This is regrettable because, as will be seen in later chapters, drilling technique can have an enormous impact on the quality of samples and in situ tests. The advice given in Chapter 7 can be used to make improvements to current specification documents. In the USA the following American Society for Testing and Materials (ASTM) standards are available for site investigation and sampling: ASTM D420—87: Standard guide for investigating and sampling soil and rock, ASTM D1452—80: Soil investigation and sampling by auger borings, ASTM D1587—83: Thin-walled tube sampling, ASTM D3550—84: Standard practice for ring-lined barrel sampling of soils, ASTM D2113—83 (reapproved 1987) Standard practice for diamond core drilling for site investigation, ASTM D4220—83: Standard practices for preserving and transporting soil samples. In the UK the provisions given in BS 5930 (which in any case describes itself as a code of practice, rather than a standard) for drilling and sampling generally are not suitable for inclusion in a contract specification (see, for example, Clayton (1986) for criticisms). 6. Groundwater. The groundwater regime is often not very well determined by ground investigation. Since pore water pressure is usually a very important factor in any engineering calculation, any seepages or inflows into the borehole should be closely monitored. Each time that groundwater is detected, the depth of entry should be measured and the speed of inflow described. Boring should be suspended and groundwater levels observed in an attempt to determine the static groundwater level. Some specifications allow for standing time (i.e. unproductive time) while the groundwater stabilizes in the borehole. Others require that the driller should only suspend work for a maximum of 20 mm. At the end of this period, if the water level is still rising, its depth is to be recorded and drilling recommenced. Each groundwater inflow should be sampled. Where water has previously been added for boring purposes, it should be bailed out before sampling. The sample should not be less than 11. 7. Storage, handling and transporting of soil samples. All samples and cores should be protected at all times from the adverse effects of weather. They should, as soon as practicable, be placed in a sample store with a humid atmosphere and a temperature between 7 and 18°C. Samples should be handled carefully at all times and should be transported to a soils laboratory for testing within two weeks of sampling. 8. In situ testing. BS 1377:1991 and ASTM Part D18 (dealing with Soils and Rocks) give specifications for the most common in situ tests, including the SPT, the cone test, vane testing and permeability testing. In addition, German standards and ISSMFE (International Society for Soil Mechanics and Foundation Engineering) Reference Test Procedures are available to cover other forms of testing (for example dynamic penetration testing). Details of these are given in Table 1.8. It is normal in the UK simply to state at the start of a Specification that all the ground investigation work is to be carried out to the British Standards for ‘Site Investigation’ (currently BS 5930:1981) and ‘Testing of Soils for Civil Engineering Purposes’ (currently BS 1377: 1991). It is a much better practice to refer specifically within the Specification to the clauses of required standard dealing with the particular test. British Standards are normally complex, and to avoid omission, specific points to be noted and adhered to by the ground investigation contractor should be highlighted within the Specification document.

29

Planning and Procurement

Test Density tests (sand replacement, water replacement, core cutter, balloon, and nuclear methods) Apparent resistivity In situ redox potential In situ California bearing ratio Standard penetration test

Table 1.8 Standards available for in situ testing British Standard American Standard BS 1377:part 9:1990, clause 2 ASTM D1556—82 ASTM D2937—83 ASTM D2167—84 ASTM D2922—91 BS 1377:part 9:1990, clause 5.1 ASTM G57—78 (re-approved 1984) BS 1377:part 9:1990, clause 5.2 BS 1377:part 9:1990, clause 4.3 ASTM D4429—84 BS 1377:part 9:1990, clause 3.3

Dynamic penetration tests Cone penetration test Vane test

BS 1377:part 9:1990, clause 3.2 BS 1377:part 9:1990, clause 3.1 BS 1377:part 9:1990, clause 4.4

Plate loading tests

BS 1377:part 9:1990, clauses 4.1,4.2

Pressuremeter test

ASTM D1586—84 ASTM D4633—86 (energy measurement) ASTM D3441—86 ASTM D2573—72 (re-approved 1978) ASTM D1194—72 (re-approved 1978) ASTM D4395—84 ASTM D4719—87

9. Journals. The information required to form the driller’s daily report must be recorded as drilling proceeds. At the end of each day’s drilling, the drilling foreman of each rig must prepare a report incorporating the following information: (i) (ii) (iii) (iv) (v) (vi) (vii) (viii) (ix) (x)

job name and location; contractor’s name; exploratory hole reference number; depth of drilling at the start and end of the shift; type of drilling rig; diameters and depths of all casing; depth of each stratum change; groundwater records; brief description of each soil type; and type, diameter and upper and lower depths of each sample, drill run, or in situ test;

for boreholes: (xi) (xii) (xiii)

locations where water was added to the boring; depths when chiselling was required; and details of instruments installed;

for drillholes: (xiv) (xv) (xvi)

orientation of the drillhole; type and diameter of barrel, and bit; and flush type, and notes on flush return and loss of return.

These records, produced on standard sheets (Fig. 1.7), should be submitted to the site engineer without fail before the start of the next day’s drilling.

30

Site Investigation

Fig. 1.7 Layout for rotary drilling daily report (from Redford 1981).

10. Laboratory testing. BS 1377:1991 and ASTM Part D18 give detailed specifications for the testing of soils, and some specifications for the testing of rocks. In addition, ISRM (International Society for Rock Mechanics) gives recommendations for methods of testing rock (Table 1.10). Table 1.9 gives details of the Specifications available at the time of writing. As with in situ testing, individual clauses should be given in the Specification, and where appropriate details requiring special attention should be highlighted.

31

Planning and Procurement

Table 1.9 Standards available for laboratory testing of soils Test British Standard American Standard Classification tests ASTM D2216—91 BS 1377:part 2:1990, clause 3 Moisture content ASTM D4643—87 ASTM D4318—84 BS 1377:part 2:1990, clauses 4,5 Atterberg limits BS 1377:part 2:1990, clause 7 Density ASTM D854—92 BS 1377:part 2:1990, clause 8 Specific gravity ASTM D422—63 BS 1377:part 2:1990, clause 9 Particle size distribution (re-approved 1972) ASTM D2217—85 ASTM D4647—87 Pinhole dispersion test Chemical tests BS 1377:part 3:1990, clause 3 Organic matter content ASTM D2974—87 BS 1377:part 3:1990, clause 4 Loss on ignition BS 1377:part 3:1990, clause 5 Sulphate content ASTM D4373—84 BS 1377:part 3:1990, clause 6 Carbonate content BS 1377:part 3:1990, clause 7 Chloride content ASTM G51—77 (reBS 1377:part 3:1990, clause 9 pH approved 1984) BS 1377:part 3:1990, clause 10 Resistivity BS 1377:part 3:1990, clause 11 Redox potential Compaction tests ASTM D698—91 BS 1377:part 4:1990, clause 3.3 Proctor/2.5kg rammer ASTM D1557—91 BS 1377:part 4:1990, clause 3.5 Heavy/4.5kg rammer BS 1377:part 4:1990, clause 3.7 Vibrating hammer Strength tests ASTM D1883—92 BS 1377:part 4:1990, clause 7 California bearing ratio BS 1377:part 7: 1990, clauses 8,9 ASTM D2850—87 Undrained triaxial shear strength BS 1377:part 8:1990, clause 7 Effective strength from the consolidated-undrained triaxial compression test with pore pressure measurement BS 1377:part 8:1990, clause 8 Effective strength from the consolidated-drained triaxial compression test with volume change measurement ASTM D3080—90 BS 1377:part 7:1990, clause 5 Residual strength by direct shear testing in the shear box BS 1377:part 7:1990, clause 6 Residual strength using the Bromhead ring shear apparatus Compressibility tests ASTM D2435—90 BS 1377:part 5:1990, clauses 3,4 One-dimensional compressibility in the oedometer Isotropic consolidation in the BS 1377:part 8:1990, clause 6 triaxial apparatus Permeability tests In the constant-head BS 1377:part 5:1990, clause 5 ASTM D2434—68 apparatus (re-approved 1974)

32

Site Investigation Table 1.10 Suggested methods for laboratory testing of rocks; ISRM Commission on Testing Methods (formerly The Commission for the Standardization of Laboratory and Field Tests) Test Reference in International Journal of Rock Mechanics Mining Science and Geomechanics Abstracts Description Petrographic description 1978, 15, (2), 41—46 Description of discontinuities 1978, 15, (6), 319—368 Index tests 1979, 16, (2), 141—156 Water content, porosity, density, absorptionrelated properties, swelling and slake durability 1985, 22, (2). 51—60 Point load strength 1978, 15, (3), 89—98 Hardness and abrasiveness 1978, 15, (2), 53—58 Sound velocity Mechanical properties 1979, 16, (2), 135— 140 Uniaxial compressive strength and deformability 1978, 15, (2), 47—52 revised 1983, Strength in triaxial compression 20, (6), 283—290 1978, 15, (3), 99—104 Tensile strength 1988, 25, (2), 71—96 Fracture toughness 1989, 26, (5), 415—426 Laboratory testing of argillaceous swelling rocks 1989, 26, (5), 427—434 Large-scale sampling and triaxial testing of jointed rock Cost considerations Most published and unpublished opinions on the methods of control and finance of site investigations in the UK express the need for more time and money (Williams and Mettam 1971; Rowe 1972), flexibility of procedure (Green 1968), and adequate liaison between geotechnical and structural design teams (Bridge and Elliott 1967). Site investigation in the UK traditionally has been carried out by specialist geotechnical contractors. These contractors vary considerably. They may be very experienced organizations controlled by qualified engineers and geologists, and supported by extensive facilities for air photograph interpretation, geotechnical laboratory testing, and computer studies: often, however, they may be organizations with limited assets, limited plant, limited engineering knowledge — and limited liability! In recent years, the financial restrictions on site investigation seem to have become tighter, but in 1972 Rowe delivered the following arguments in favour of spending more on site investigation. 1. It is known that more claims by piling contractors arise due to poorly or inaccurately known ground conditions than to any other cause (Tomlinson and Meigh 1971). 2. Site investigation costs are very low compared with the cost of earthworks or foundation construction, and even smaller as a proportion of the total capital cost of the works, can be seen in Table 1.11. These figures represent a decline in expenditure since the 1940s since Harding (1949) reckoned the cost of site investigations for ‘fair-sized works’ to be usually about 1 to 2% of the cost of the main work.

33

Planning and Procurement The influence of incorrect site investigation data on the final cost of a project is difficult to assess but can be very large. Rowe cites examples of a case where the omission or inclusion of sand drains could make a difference of 2 to 5% of total project cost, and where unnecessary foundation treatment added 5% to the cost of the works. These figures are certainly not representative of the upper end of the spectrum, as claims for unforeseen ground conditions can easily amount to 10% of contract value. Table 1.11 Site investigation costs (from Rowe (1972)) % of capital % of earthworks and Type of work cost of works foundation cost Earth dams 0.89—3.30 1.14—5.20 Embankments 0.12—0.19 0.16—0.20 Docks 0.23—0.50 0.42—1.67 Bridges 0.12—0.50 0.26—1.30 Buildings 0.05—0.22 0.50—2.00 Roads 0.20—1.55 (1.60)?—5.67 Railways 0.60—2.00 3.5 Overall mean 0.7 1.5 As noted earlier, Tyrrell et al. (1983) found, in an analysis of ten selected highway contracts, that additional expenditure rose to an average of 35% of the tender value. Of this about one-half could be attributed to geotechnical matters. On this basis it is easy to argue for an increase in expenditure on site investigation. But it has proved difficult to establish that increased expenditure on site investigation leads to reductions in construction cost over-runs. What is required is that all expenditure on ground investigation sitework and testing, which typically amounts to 60—70% of the total cost of a site investigation, should be carefully targeted at giving information required for particular and welldefined geotechnical problems. This will lead to reductions in expenditure in some cases, and increases in others. PROCUREMENT

In the UK it has been widely considered that procurement, in its broadest sense, is the key to obtaining a good site investigation at a reasonable price. Investigations carried out for the Construction Industry Research and Information Association (CIRIA) have been reported by Uff and Clayton (1986, 1991). Many of their more general recommendations are incorporated into the earlier parts of this chapter; only those dealing with the detailed mechanisms of procurement are considered below. The way in which ground investigation work is organized has been described briefly above, under ‘Implementation’ and ‘Planning’. A number of different organizational models are used worldwide, as can be seen from the examples given in Fig. 1.8. In essence, a good system of procurement will ensure that key elements of work are carried out properly, by competent personnel. In Fig. 1.8, examples A, B and C are satisfactory; D and E omit significant parts of the investigation process, and are bad. It is essential that: 1. desk study, air photo interpretation and a walk-over survey are carried out; 2. the ground investigation is properly designed, taking into account the probable ground conditions and the proposed construction; 3. the required type and standards of ground investigation field and laboratory work are properly defined; 4. during the ground investigation, standards are enforced by competent supervision; and 5. as ground investigation proceeds, the ground conditions are reassessed in the light of information emerging from the work, and that work is rescheduled if necessary.

34

Site Investigation

Fig. 1.8 Examples of division of site investigation work. Procurement methods will often concentrate upon obtaining minimum prices, without considering how well the required quality of ground investigation work can be defined. This is a serious mistake, since many of the activities involved cannot very readily be checked. For example, a good quality standard penetration test requires attention not only to the test equipment and the method used for the test itself, but also to the method of boring, and the water levels within the borehole, both before and during the test. The end product is a series of numbers, the validity of which can be known only if all these matters have been observed, reported, and considered. Therefore, it is suggested that the procurement system should aim: 1. to ensure that a competent geotechnical adviser is retained by the promoter/ developer at an early stage during the conceptualization of the project, in order to guide the project; 2. as far as possible, to give overall responsibility for all geotechnical matters to a single individual or company; and 3. to select geotechnical advisers and contractors on the basis of their resources (staff, equipment, etc.), and experience with similar forms of construction and ground conditions, and not primarily on the basis of their fee level or unit rates. In the UK, two systems of procurement of site investigation are in common use, as detailed in CIRIA Special Publication SP45 (Uff and Clayton 1986). System 1: Use of a geotechnical adviser with the separate employment of a contractor for physical work, testing, and reporting as required In this system the desk study, the planning and supervision of any fieldwork (such as boring, drilling, trial pitting or in situ testing) and laboratory testing work that may be necessary is carried out by the

35

Planning and Procurement geotechnical adviser. He will often be a member of a firm of civil engineering consultants but may also be a specialist geotechnical consultant. This system is widely used on large civil engineering projects. The geotechnical adviser will normally be employed by the developer under the Association of Consulting Engineers Conditions of Engagement, while the specialist ground investigation contractor will be chosen by competitive tender and will work under ICE Conditions of Contract. Two versions of ICE contract are in use; the ICE 5th Edition and the ICE Conditions of Contract for Ground Investigation. When using this system it is important that the developer or his advisers should check that the chosen geotechnical adviser has sufficient geotechnical skill to carry out the desk study, plan and supervise the ground investigation and interpret its results. It is possible to make use of the contractor’s engineering skills only after the tendering process. Therefore the skills of the geotechnical adviser are extremely important. The geotechnical adviser is expected to carry out a thorough desk study and plan an investigation appropriate to the needs of the developer. This is then used to prepare a specification and bill of quantities which, together with the conditions of contract, form the basis of the tender for the field and laboratory work to be carried out by a specialist contractor. Generally between three and four companies should be selected by the geotechnical adviser to tender for the field and laboratory work, on the basis of their previous experience of this type of work, the skills of their staff and the amount and quality of their equipment. The lowest submitted tender price is generally accepted but the contract is subject to remeasurement as the work proceeds. The final cost to the developer of the entire ground investigation will be the sum of the final contract price after measurement and the professional fees paid to the consulting engineer. This system has been found to work well provided that: 1. an adviser with a sufficient number of skilled geotechnical staff is engaged; 2. a thorough desk study, made by the geotechnical adviser, is used as the basis for the planning of any programme of drilling and testing; 3. not more than four specialist contractors arc asked to tender and the selection of these companies is rationally and thoroughly carried out; and 4. proper levels of supervision are provided by the geotechnical adviser in the control of field and laboratory work. Supervision is the key to the successful use of System 1. In certain cases it may be advantageous for parts of the work to be done by the contractor on a dayworks basis. Under System 1, the work to be carried out by the contractor must be closely defined before the contract is let and must be paid for at fixed rates independent of the time taken to carry it out. If the work is particularly important to the success of the investigation, if it is very complex, or if the geotechnical adviser needs to be able to vary the work as it proceeds, dayworks payments may be helpful. For example, dayworks could be used to pay for plate loading tests, for drilling and boring in key zones, or for time spent in investigating groundwater conditions. It is also possible to pay a specialist contractor to carry out the reporting of an investigation; this is better done on an hourly basis rather than by lump sums. System 1 has the advantage of using forms of contract that are well known in the civil engineering construction field and it can be used to demonstrate cost-accountability through the tendering process. This is the most commonly used form of procurement for larger ground investigations and is therefore well understood. Its difficulties lie in the complexity of its contractual arrangements, the need to ensure that sufficient expertise and supervision are provided by the geotechnical adviser and the division of responsibility for the satisfactory outcome of the investigation between the geotechnical adviser and the contractor. It has frequently been said that the method of competitive tendering commonly associated with this system, and the consequent low prices paid to contractors for investigation work, is a major cause of low-quality investigation. This problem, however, is a consequence of too large tender lists prepared without detailed selection of tenderers. It is not necessarily a result of using the system.

36

Site Investigation

System 2: Package deal contract, with desk study, planning and execution of field and laboratory work, and reporting, being carried out by one company or a consortium No formal conditions of contract exist for this system, although draft documents have been proposed in CIRIA Special Publication 45. Despite the lack of published conditions of contract, versions of this system are in common use to obtain ground investigations for low-rise building development. The system is also used for large site investigation contracts carried out abroad, for example in the Middle East. In this system the developer selects up to three specialist ground investigation companies on the basis of past experience, reputation, and published information relating to specialists in the field. Information on companies and individuals is available from: • • • •

the British Geotechnical Society; the Association of Ground Investigation Specialists; the Institution of Civil Engineers; and the Geological Society.

The companies selected may be either ‘contractors’ or ‘consultants’ according to the British Geotechnical Society’s Directory, but they should have sufficient qualified and experienced staff to be able to carry out the proposed size of investigation. On the basis of a preliminary desk study, the companies offer to carry out a complete site investigation, including desk study, air photograph interpretation, design and execution of ground investigation and reporting, either for a lump sum or on the basis of measurement of work agreed as the investigation progresses. The specialist company that carries out the work is expected to supervise its own drilling and testing and will be liable under the 1982 Supply of Goods and Services Act both for the quality of work and for any recommendations that are made in the report of the investigation. The advantages of System 2 to a developer are that a lump sum contract can be negotiated; this is obviously important when carrying out financial forecasting. A further advantage is that the responsibility for ground investigation is not divided between two parties, as in System 1. Because of the cost to the tenderers of preliminary desk studies, it is unlikely that lump sum contracts can be used for very large civil engineering projects, but this type of procurement will certainly be more suited than System 1 to many low-rise building developments, because of its relatively simple contract documentation and its flexibility. An advantage of this system is that the leading design professional (who might typically be an architect in the case of a low-rise building development) is not necessarily required to have geotechnical skill and experience of ground investigation techniques. If he does not possess such skill, however, it becomes extremely important that care is taken in the selection of ground investigation specialists who are suitable for the complexity of work to be carried out. A possible disadvantage of System 2 is the lack of well-tried and proven contract documentation. However, this does not appear to have prevented the successful use of this method of procurement in recent years. To overcome this it is suggested that the contract documents used are those given in the appendices to CIRIA Special Publication 45. EXECUTION

Supervision A good site investigation is made in the field. Engineering excellence, sophisticated laboratory

37

Planning and Procurement techniques and the use of powerful computational methods cannot ever be expected to make any contribution to a site investigation performed by bad drillers without engineering supervision. Since this often occurs, it is hardly surprising that the value of site investigation is sometimes questioned by engineers not familiar with its techniques. Supervision of any site investigation requires an engineer who is familiar with: 1. the techniques of investigation; and 2. the objectives of the particular investigation. This engineer is the key person in ensuring that the best use is made of the expenditure on site investigation, and to this end he must spend a very large part of his time on site during the investigation. While on site, the supervising engineer should: 1. closely watch the drilling and sampling techniques, to ensure that disturbance of soil is minimized; and the techniques and equipment comply with the specification; 2. frequently check the records of borings provided by the drillers for authenticity and accuracy; 3. carry out sample description and prepare engineering logs, except on small investigations where it will be more economical to transport samples to a laboratory for description; 4. liaise with the structural design engineer, so that the investigation can be modified as a result of its initial findings; 5. ensure that driller’s and engineer’s borehole records, and the samples are despatched to the soil laboratory at frequent intervals; 6. provide conditions of storage for the samples on site which ‘will not lead to their deterioration; and 7. check the adequacy of sample sealing by the rig foremen. Quite clearly, it will be difficult for one man to supervise more than one drilling rig satisfactorily, and for this reason drilling technicians have sometimes been used. A drilling technician will be assigned to one rig, and will be responsible for all the technical aspects of that rig’s work. He would normally, for example, prepare the records of drilling, instruct the driller at which level to take samples, and carry out in situ testing and the installation of instrumentation. In the majority of cases, both in the UK and overseas, drilling technicians are not used. Site investigations are usually carried out by drillers who do not understand the mechanisms of disturbance of soil samples, who are not informed of the objectives of the individual investigation, and who are often motivated solely by productivity bonuses. Under these conditions, the supervising engineer is the only force available in the struggle to produce a sound investigation. To be effective, the supervising engineer must understand the practical aspects of drilling. The key points in checking the effectiveness of a site investigation are as follows. 1. Avoid excessive disturbance. Look for damaged cutting shoes, rusty, rough or dirty sample barrels, or badly designed samplers. Check the depth of casings to ensure that these never penetrate beneath the bottom of the borehole. Try to assess the amount of displacement occurring beneath power augers, and prevent their use if necessary. 2. Check for water. Ensure that adequate water levels are maintained when drilling in granular soils or soft alluvium beneath the water table. The addition of water in small quantities should be kept to a minimum, since this allows swelling without going any way towards replacing total stress levels. Make sure the driller stops drilling when groundwater is met. 3. Check depths. The depths of samples can be found approximately by noting the number of rods placed on the sampling tool as it is lowered down the hole, and the amount of ‘stick-up’ of the last rod at the top of the hole. This type of approach is often used by drillers, but is not always satisfactory. Immediately before any sample is taken or in situ test performed the depth

38

Site Investigation of the bottom of the hole should be measured, using a weighted tape. If this depth is different from the last depth of the drilling tools then either the sides of the hole are collapsing, or soil is piping or heaving into the base. Open-drive sampling should not then be used. 4. Look for faulty equipment. On-site maintenance may lead to SPT hammers becoming nonstandard, for example owing to threading snapping and the central stem being shortened, giving a short drop. When working overseas with subcontract rigs the weight of the SPT hammer should also be measured. Other problems which often occur are: (i) the blocking of vents in sampler heads; and (ii) the jamming of inner barrels in double tube swivel-type corebarrels. 5. Examine driller’s records regularly. The driller should be aware that the engineer is seeking high quality workmanship. One of the easiest ways of improving site investigation is to demand that up to the moment records are kept by the driller as drilling proceeds. These should then be checked several times a day when the engineer visits the borehole. Any problems encountered by the driller can then be discussed, and decisions taken.

Safety Safety should be of major concern during the fieldwork and laboratory testing phases of ground investigation. Potential hazards include: 1. incapacity as a result of prolonged exposure to bad working conditions (for example, deafness as a result of exposure to high levels of noise); 2. injury or death as a result of misuse of plant and equipment (for example, using frayed winch ropes, not setting up drilling equipment in a stable configuration, etc.); 3. injury or death as a result of contact with overhead electricity cables (particularly by contact with drilling rig masts, but also with cranes during transporting); 4. injury or death as a result of excavation through services (electricity, gas, water, etc.), during boring, drilling or pitting; 5. injury or death as a result of explosions of gases emanating from the ground (for example methane from landfill); 6. injury or death as a result of collapse of trenches on to personnel carrying out logging or sampling; 7. damage to health as a result of contact with contaminated ground, or in the laboratory, working with contaminated samples; 8. poisoning, as a result of inhaling or ingesting toxic gases or substances such as asbestos, cyanides, etc; 9. damage to health, or death, as a result of radiation; and 10. damage to health, or death, as a result of contact with animal carcases or sewage (leading, for example, to anthrax or Weil’s disease). Whilst many of these risks are associated with the investigation of contaminated land, a very large proportion may be present during any site investigation. Under the UK ‘Health and Safety at Work etc.’ Act, all persons involved with an investigation have a responsibility to see that safe working practices are adopted. This includes the promoter of the development, who may have knowledge of previous site use, the consulting engineer, who must ensure that sufficient resources are devoted to a safety assessment before field and testing work is specified, and the specialist ground investigation contractor, who must enforce safe working practices during the ground investigation. Engineers and geologists will be particularly responsible, since they will be directly in control of those most exposed to risk. The construction industry has a poor safety record, and there is always the temptation to reduce costs by taking short cuts with safety. This must be prevented. To help in the drive for greater safety in ground investigation the British Drilling Association have recently published two reports:

39

Planning and Procurement

Code of Safe Drilling Practice, British Drilling Association, Brentwood, Essex, UK (March 1992). Guidance Notes for the Safe Drilling of Landfills and Contaminated Land, British Drilling Association, Brentwood, Essex, UK (March 1992). It is recommended that all professionals involved in ground investigation should study both these and the literature to which they refer, before carrying out fieldwork.

Quality assurance Quality assurance is ‘All those planned and systematic actions necessary to provide adequate confidence that a product or service will satisfy given requirements of quality’. In other words, quality assurance concerns the management of an organization to meet agreed quality objectives. In itself it does not guarantee that a service is of the necessary quality for a given job, but attempts to meet predetermined standards by approaching the work in a systematic manner. In this sense it simply represents good management practice. In the UK, quality systems are now being implemented in the ground investigation industry. They are standardized internationally (ISO 9001—1987), in Europe by CEN (EN 29001—1987), and in the UK (BS 5750:1987) under identical documents. Quality systems comprise several levels of activity (Fig. 1.9): 1. Quality policy. The overall quality intentions and direction of an organization as regards quality, as formally expressed by top management. 2. Quality management. That aspect of the overall management function that determines and implements the quality policy. 3. Quality system. The organizational structure, responsibilities, procedures, processes and resources for implementing quality management. 4. Quality control. The operational techniques and activities that are used to fulfil requirements for quality.

Fig. 1.9 Relationships of quality concepts (BS 5750: part 0: section 0.1: 1987). In the UK, ground investigation industry quality assurance is being applied at two levels. First, ‘internal quality assurance’, which aims to provide the management of an organization with the

40

Site Investigation confidence that the intended quality is being achieved, is being implemented. Under BS 5750, quality systems can be audited by a third party ‘certification’ body, such as Lloyds or the British Standards Institution. Since ground investigation generally has a rather short duration, it is sensible that at the outset it is the supplier who leads the quality process — it has been found that attempts to impose ‘external quality assurance’, i.e. activities which aim to provide confidence that the supplier’s quality system will provide a product that will satisfy the client’s stated quality requirements, are difficult to set up in the absence of legislation, for such small and diverse projects. Secondly, laboratory testing services are becoming subject to third-party accreditation by the National Measurement Accreditation Service (NAMAS). This represents less of a problem, in theory, because British Standards provide tight specifications for most aspects of the more commonly used tests. At the time of writing (1992) the UK’s Department of Transport has stated that all work carried out on UK highway investigation after April 1993 must be carried out in NAMAS accredited laboratories. Presently there is only one major geotechnical laboratory accredited for a wide range of soil tests, although it is expected that several will be able to offer an accredited service by the end of 1992. In procuring the services of geotechnical specialists, whether consultants, contractors, or specialist sub-contractors, it is recommended that, other things being equal, those who offer a certified quality management system, or an accredited laboratory or field testing service should be favoured.

41

Chapter 2

Description and classification of soils and rocks

INTRODUCTION

From an engineering viewpoint, the ground beneath a site can conveniently be divided into the categories shown in Table 2.1, which are based upon generalizations of its expected behaviour in construction works. These broad generalizations are, of course, limited in accuracy. But they give the geotechnical engineer a good basis on which to consider, at the start of a project, both the likely construction problems and the methods of investigation which might be used. In practice, it is found that the ground varies continuously beneath a site, and it is not often possible to find sharp transitions from one type of material to another. This then, calls for more refined, systematic, description and classification of soils and rocks. The history of soil classification and description has been described by Child (1986). Early attempts to divide soil into different categories used laboratory testing, typically particle size distribution. Casagrande (1947) considered that much of this ‘textural soil classification’ was unreliable, because it did not always reflect the effects of silt and clay on the engineering behaviour of the mass. On the basis of his experience he considered that only the Airfield Classification System, which was based on the results of particle size analysis and plasticity (Atterberg limit) tests (see Chapter 8), was suitable for an assessment of soil for airfield pavement design. It was noted that, with experience, soil might be classified on the basis of visual/manual description alone; Casagrande therefore suggested what he termed ‘descriptive soil classification’. American practice therefore developed in two directions. Soil Classification, based on Casagrande’s Airfield Classification System, became standardized (ASTM D2487— 92). As originally proposed, the system is based solely on particle size distribution and plasticity tests, and soils are designated by letters alone, depending upon which group they fall into. At the same time, a soil description system has been developed, based upon visual examination and simple land tests (ASTM D2488—69). Table 2.1 Categories of ground beneath a site Material type Strength Compressibility Permeability Rock Very high Very low Medium to high Granular soil High Low High Cohesive soil Low High Very low Organic soil Very low Very high High Made ground Medium to very low Medium to very high Low to high Early British practice was summarized in Cooling et al.’s discussion of Casagrande’s 1947 paper. Almost without change, this appeared in the first British Code of Practice on Site Investigations, Code of Practice No. 1 (1950). It was republished as Table 1 of CP 2001:1957. Soil description was based on estimated mass engineering behaviour, and used simple visual description and hand tests. A classification system based on Casagrande’s work was included for roads and airfields work, that is to assess the behaviour of materials during compaction and under pavements. Subsequent development 1

Description and Classification of Soils and Rocks occurred primarily through work by Akroyd (1957), Dumbleton (1968, 1981), Dumbleton and Nixon (1973), and the Engineering Group Working Party on ‘The Preparation of Maps and Plans in Terms of Engineering Geology’ (Geological Society of London 1972). BS 5930:1981 contained two parts, a soil description based on mass behaviour, and the British Soil Classification System (BSCS) based on the work of Dumbleton (1968, 1981). SOIL AND ROCK DESCRIPTION Soil and rock description is to a certain degree subjective. In order to minimise the subjective element a systematic examination should be carried out using a standard terminology, whether the material be in a natural exposure, trial pit face or samples recovered from a borehole. The use of a standardised scheme of description ensures that: (i) all factors are considered and examined in logical sequence (ii) no essential information is omitted (iii) no matter who describes the sample, the same basic description is given using all terms in an identical way (iv) the description conveys an accurate mental image to the readers (v) any potential user can quickly extract the relevant information. Norbury et al., 1986

The engineering description of the ground conditions beneath a site is a progressive exercise which at each step involves further departure from strictly factual description, and thus an increased interpretative element. Three steps are involved. 1. The description of individual samples from a borehole, each sample being described in isolation and in completely factual terms, noting any disturbance or obvious loss of material caused by sampling. Any two geologists or engineers with sufficient and comparable experience should produce almost identical descriptions. 2. The combination of these individual descriptions to form a stratum description on the borehole log. In so doing, the engineer or geologist will take into account the information on the ground conditions, depths to strata changes, groundwater levels, field and laboratory test results, etc., given on the driller’s daily record sheets. Interpretation is necessary, and so, as Norbury et al. have stated ‘strictly, there is no such thing as a “factual” borehole log’. 3. The drawing together of individual borehole, trial pit and exposure records, to arrive at an assessment of the mass properties of the various strata, their geometric distribution, and their variability, in a summary in the text of the ground investigation report. In this process the skills of soil and rock description, coupled with experience, are paramount. SOIL DESCRIPTION

The description of soil is currently covered, in the UK, by BS 5930:1981. At the time of writing BS 5930 is under revision. The system of soil description given below therefore does not follow completely the code, but takes into account both current good practice, and changes which have been proposed. Samples must be described in a routine way, with each element of the description having a fixed position within the overall description: a) consistency or relative density; b) fabric or fissuring;

2

Site Investigation c) d) e) f) g) h) i)

colour; subsidiary constituents; angularity or grading of principal soil type; PRINCIPAL SOIL TYPE (in capitals); more detailed comments on constituents or fabric; (geological origin, if known) (in brackets); and soil classification symbols (optional).

Descriptions should be simple, since very detailed comments on all aspects of a soil lead to confusion. Some examples are given below: Very stiff fissured dark grey CLAY (London clay) (a) (b) (c) (f) (h) Loose brown very sandy subangular coarse GRAVEL with pockets of soft grey clay (a) (c) (d) (e) (f) (g) Firm laminated brown SILT and CLAY (a) (b) (c) (f) (f)

Soil types In routine soil description, the material being considered is first placed into one of the principal soil types in Table 2.2. Table 2.2 Principal soil types Soil type Description Cohesive soil CLAY Granular soils BOULDERS Granular soils COBBLES Granular soils GRAVEL Granular soils SAND Granular soils SILT Organic soil PEAT Man-made soils and MADE GROUND other materials Most soils will be composed of a variety of different particle sizes, some of which may be cohesive. Whilst classification is concerned only to determine the proportion by weight of each constituent, description is carried out to ascertain probable engineering behaviour. In this sense, we adhere to the proposals of CP 2001 (1957) and reject BS 5930:1981, following Norbury et al. (1986) (Table 2.3). CP 2001

Table 2.3 Comparison of CP 2001 and BS 5930 BS 5930

Soils possessing cohesion and plasticity are described as fine soils, although the majority of the soil by weight may be coarse or very coarse soil. It is not possible to give a percentage of clay and/or silt above which they become the principal component, since the mass behaviour depends on the mineralogy of the soil particles. The description is based on engineering judgement.

Soils with more than 35% clay and/or silt are described as either clay or silt. Soils with less than 35% are described in terms of coarse or very coarse soils, irrespective of whether they have cohesion and plasticity. The description is therefore based on the particle size distribution, but the division between silt and clay is strictly on the Atterberg limits. These factors can be difficult to assess visually for some materials and laboratory tests are required to confirm descriptions.

3

Description and Classification of Soils and Rocks

The fundamental difference between these two proposals is that under the BS 5930:1981 proposals, many ‘clays’ i.e. materials of low strength, very low permeability and high compressibility, must be termed silts, sands or gravels. In practice the addition of about only 12% by weight of clay is required to make a well-graded granular material perform in engineering works as a cohesive soil. Therefore the stance of BS 5930:1981 is considered untenable. The first stage of the description process is the identification of the principal soil type, on the basis of the expected behaviour of the soil mass. In soils where the granular fraction dominates behaviour (termed ‘granular soils’), the principal soil type is identified on the basis of a particle size. As an aid to visual identification, it should be noted that coarse silt represents the normal limit of resolution of individual grains with the unaided eye. The principal soil type is the (single) component of the soil (i.e. boulders, cobbles, gravel, sand or silt) which is thought to represent (in a coarse soil) the greatest proportion by weight (Table 2.4). Where two types are thought to be equal, two components may be given (for example, sand and gravel, boulders and cobbles). BS 5930:1981 rightly noted that the properties of very coarse soils (i.e. boulders and cobbles) cannot be reliably estimated from boreholes-trial pits or exposures must be used. Table 2.4 Identification of principal soil type Principal soil type Particle size (mm) Particle size (mm) Boulders >200 Cobbles 60—200 coarse 20—60 Gravel 2—60 medium 6—20 fine 2—6 coarse 0.6—2 Sand 0.06—2 medium 0.2—0.6 fine 0.06—0.2 Silt 75% organic matter); 2. describe degree of humification in accordance with the von Post method (Table 2.5); and 3. where possible, give basic fibre details. Examples of description would be: • •

black amorphous PEAT (H6); brown non-woody fine fibrous PEAT (H2).

It is extremely important to attempt, during the description of near-surface soils, to identify made ground. Such material may be: • • •

compressible; highly variable; chemically contaminated.

Made ground is ground filled by man’s activities, rather than as a result of geomorphological processes. Made ground may comprise (for example) compacted granular fill, in which case it may be extremely difficult to detect from the description of an isolated sample. At the other extreme, it may result from the tipping of household waste. In either case, sample description can only hope to identify made ground by searching for man-made artefacts, such as fragments of brick, clinker, tile, glass, etc., and at the other end of the scale, concrete, cars and parts of machinery, paper and plastics. Table 2.5 Assessment of degree of humification (after von Post (1922)) Degree of humification

Decomposition

H1

None

H2

Insignificant

H3

Very slight

H4

Slight

H5

Moderate

H6

Moderately strong

Plant structure Easily identified Easily identified Still identifiable Not easily identified Recognizable, but vague Indistinct (more distinct after squeezing)

H7

Strong

Faintly recognizable

H8

Very strong

Very indistinct

H9

Nearly complete

Almost not recognizable

H10

Complete

Not discernible

Content of amorphous material

Material extruded on squeezing (passing between fingers)

None

Clear, colourless water

None

Yellowish water

Slight Some Considerable

Brown, muddy water; no peat Dark brown, muddy water; no peat Muddy water and some peat

Considerable

About one-third of peat squeezed out; water dark brown

High

About one-half of peat squeezed out; any water very dark brown

High

Nature of residue

Not pasty Somewhat pasty Strongly pasty

Fibres and roots more resistant to decomposition

About two-thirds of peat squeezed out; also some pasty water Nearly all the peat squeezed out as a fairly uniform paste All the peat passes between the fingers; no free water visible

5

Description and Classification of Soils and Rocks

As with very coarse soils, made ground is best described in an exposure or a trial pit. But in this case there will be much greater safety considerations. Made ground can often produce poisonous gas (as a result of decomposition of organic material), will contain sharp materials, likely to injure (glass, metals, etc.), and may give rise to instability in the sides of trial pits. It is not advisable to allow work to be carried out from within trial pits in made ground unless proper support, breathing apparatus, and full protective clothing are available. There are no formal systems in use for the description of made ground. Where the made ground resembles natural soil, then normal soil descriptions can be used, but with additional comments added to inform the reader as to how the material has been identified as made ground (for example, ‘scattered brick and tile fragments’). In all cases the following should be noted: • • • • • • • •

organic matter, and its degree of decomposition; smell; striking colours; signs of heat (combustion); presence or absence of large objects (concrete blocks, cars, fridges, etc.) voids, hollow objects; other compressible materials; and anything by which the made ground may be dated (for example, product labels, old newspapers).

Secondary constituents Where soils are composed of a variety of different constituents, a simple scheme is required to allow them to be identified. Norbury et al. (1986) have criticized the complexity of the proposals in BS 5930, and there is certainly considerable evidence that they have not been applied in practice. It is suggested, following Norbury et al., that the scheme of Table 2.6 should be adopted.

Principal soil type Granular soil types (boulders, cobbles, gravel, sand, silt) Cohesive soil types (clay)

Table 2.6 Identification of soils Terminology used Approx. % of secondary Before principal After principal constituent constituent constituent (by weight) 0 — — 50 Very dense However, different components have used different systems. In some the density descriptor (for example, medium dense) used on a borehole log was obtained simply by averaging the SPT in a given stratum, and looking up the appropriate term in Table 2.7. In others, following Gibbs and Holtz (1957) work on the effect of overburden presence on the SPT ‘N’ value, ‘N’ values were corrected to a common overburden pressure before the descriptor was chosen. These two procedures will produce quite different results either at very shallow depth or at very great depths. Following Skempton (1986) and Clayton (1992) the procedure proposed is as follows. 1. ‘N’ values should be corrected for effective overburden pressure and energy (see Chapter 10), to give (N1)60 2. In coarse granular soils, it should be noted that the N value will be high for the in situ relative density. There is now increasing evidence that coarse (i.e. gravel) particles significantly increase penetration resistance, and the classification given below, which was derived for sands, will overestimate the density of gravels. 3. The density descriptor should be selected from Table 2.8. Table 2.8 Selection of density descriptor (N1)60 (blows/300 mm) Density descriptor 0—3 Very loose 3—8 Loose 8—25 Medium dense 25—42 Dense 42—58 Very dense During sample description, simple hand tests are used to describe the consistency of cohesive soils. Consistency is the estimated undrained shear strength of the intact blocks of a soil. Large diameter triaxial tests carried out on fissured clays will normally give much lower values of undrained shear strength because of the weakening effect of the fissures. Three simple tests are commonly used to determine consistency. These use the hand and fingers, a pocket penetrometer or a hand vane tester (Fig. 2.2). The vane test will normally give undrained shear strengths in engineering units, but most pocket penetrometers are calibrated in terms of unconfined compressive strength; this should be halved to find the undrained shear strength. All tests used to find consistency are carried out on small

8

Site Investigation samples at a rate much faster than is used in laboratory strength testing. For these reasons their results should not normally be used in engineering calculations. Some penetrometers have in-built (empirical) corrections, which are intended to correct for size and rate effects. These corrections must be removed by calculation before the strength descriptor is selected from the list below.

Fig. 2.2 Hand vane and pocket penetrometer. Where strength measurements cannot be made, the consistency of a soil can be determined with reasonable accuracy using hand and fingers, as follows: 1. 2. 3. 4. 5. 6.

Very soft Exudes between fingers when squeezed in hand Soft Moulded by light finger pressure Firm Can be moulded by strong finger pressure Stiff Cannot be moulded by fingers Can be indented by thumb Very stiff Can be indented by thumb nail Hard Cannot be indented by thumb nail (not included in BS 5930:1981).

Where a sample straddles a boundary, this can be indicated as, for example, ‘firm to stiff’. On the engineer’s borehole record, where a gradual or sharp change in consistency occurs within a clay formation, it is common to use a description such as ‘firm, becoming stiff from 12 m, etc.’. Alternatively, if the formation has been fully described before the depth has been reached at which a change in consistency occurs, one can put ‘...becoming stiff at 12 m’ opposite the relevant depth. Consistency is related directly to strength as shown in Table 2.9. The blows necessary to drive open-drive sampler tubes such as the SPT split spoon or U100 into the ground are sometimes used in estimating consistency. Opinions on the use of the SPT N value to determine undrained shear strength vary. De Mello (1969) shows that correlations between strength and N value are generally poor, but Stroud (1974) has found good correlations for heavily overconsolidated soils, mainly in the UK (Fig. 9.3). There seems little point in attempting to correlate 9

Description and Classification of Soils and Rocks the blows required to drive a U100 sampler with soil strength, because the weight of the hammer and ancillary equipment can vary considerably. Some drilling companies use a jarring link and sinker bar at the base of the hole, while others rod up to the top of the hole and drive the tube with an SPT trip hammer. Table 2.9 Relationship between consistency and undrained shear strength of intact soil Undrained shear strength Consistency p.s.i. p.s.f. t.s.f. kN/m2 Very soft 90

CV CE

Peat soils consist predominantly of plant remains which may be fibrous or amorphous.

Organic matter suspected to be a significant constituent. Example MHO. Organic SILT of high plasticity.

of high plasticity 50 to 70

CH

Pt

C

of intermediate plasticity 35 to 50

CI 65 to 100

CLAY of low plasticity 60 Boulders and cobbles Coarse-grained 2—60 Gravel Medium-grained 0.06—2 Sand Fine-grained 0.002—0.06 Silt Very fine-grained 80l00m. Disadvantages: heavy and cumbersome and require large power supply. 1000-12000 0.2-1 Depth resolution excellent, but Pinger Shallow penetration varies in the inverse reflection ratio to the sea bed sediment grain (penetration size. through clayey soil only) Side scan Sea floor 48000-105000 Resolution good, equipment sonar profiling compact and requires small power supply. Conventional Sea floor 30 000-40 000 As above. echo profiling sounding Air gun Deep and 2-25000 1-4 Capable of deep penetration in shallow shallow to moderate depths of reflection water. Moderate penetration in deep water. * Precision boomer.

The success of both land-based and marine seismic reflection methods depends on the velocity contrast between different materials. If the velocity contrast is low, then most of the energy will be refracted, and hence the reflected energy will be small. In general only about 1% of the incident energy is reflected back, except in exceptional circumstances, such as reflections from the sea floor where the reflection coefficient is normally very high. One advantage of reflection methods over refraction methods is that interfaces with low velocity contrasts arc more readily identifiable by increasing the amount of energy put into the ground. This may, however, create problems with unwanted multiple reflections from strong reflectors. Seismic reflection methods are used mainly to profile subsurface or sub-bottom (in marine investigations) layered systems. Current research, however, shows that the data acquired by

43

Site Investigation such methods are capable of a more detailed interpretation to obtain geotechnical parameters. McCann and Taylor-Smith (1973) review the application of seismic reflection together with other geophysical techniques in the determination of geotechnical parameters of sea-floor sediments.

Seismic tomography Tomography (from the Greek, tomos: a slice or section) is a method whereby an image of the distribution of a physical property within an enclosed region is produced without direct access to the zone. Seismic velocity tomography is a geophysical sectioning technique which determines the spatial seismic velocity distribution within a given area of interest. Seismic velocity tomography is potentially extremely useful in geotechnical ground investigations, for two reasons. First, tomography can give information on the general variability (i.e. of seismic velocity) beneath a site, and by inference allow a qualitative assessment of the variability of properties such as stiffness and strength. Features such as voids, fractures, rock layers and soft spots are often difficult to detect using more conventional (boring and drilling) techniques, because only a minute proportion of the subsoil is sampled and tested by direct methods of ground investigation. Geophysical techniques may be helpful in detecting such features. Secondly, the technique can be used to provide values of the ‘very small strain’ stiffness (Go, or Gmax) of the ground, since this is uniquely linked to seismic shear wave velocity (Vs) through the equation: 2 G = Vs ρ (4.10) where = bulk density of the soil. Depending upon the ground conditions and the strain levels imposed by construction, this stiffness may be more or less relevant to the operational stiffness required for the analysis of geotechnical displacement problems. In many ground conditions (for example fractured rock or granular soils) the alternative methods of estimating stiffness may be sufficiently expensive or inaccurate as to make seismic methods attractive. The form of tomography that is, perhaps, most familiar is the technique of computer assisted tomography (CAT), as used in diagnostic medicine. The method of seismic tomography has been in use for some years: amongst the first applications of geotomography was a survey by Bois et al. (1971) in an oil field. In seismic velocity tomography, an image of the distribution of the seismic propagation velocity properties within a region of the ground is deduced from measurements of the transit times of artificially induced seismic waves crossing the zone. The process can be divided into a number of activities, each of which requires careful attention if useful results are to be achieved: • • •

data acquisition and reduction; reconstruction of velocities; and interpretation.

The field geophysical method involves the input of seismic energy (from mechanical hammers, or sparkers) at discrete points down boreholes, and possibly along the ground surface, and the acquisition of seismic records of the incoming wave energy (via geophones or hydrophones) at as many other discrete points around the area of interest as is feasible. The seismic traces are then inspected to determine the first arrival of the wave type of interest. In many rock investigations this may be the onset of primary, compressional (P) waves, but in weaker sediments such as saturated soils, it may be necessary to determine the arrival time of secondary, shear (S) waves. The data required from the field, before reconstruction processing can commence, are the travel times from each source to every receiver, and the co-ordinates of the source and receiver positions.

44

Subsurface Exploration: Engineering Geophysics Tomographic reconstruction (as applied here) is the mathematical process by which velocities at different points within the ground are calculated. The principles of tomographic reconstruction are well-established (Radon 1917). In general, the region to be imaged is divided into discrete rectangular areas or cells (Fig. 4.25). Let j be the seismic propagation velocity within the jth of n reconstruction cells. This velocity value applies uniformly across the full area of the cell. The travel time, ti for the ith ray across the grid of cells is given by: n d ij ti = (4.11) j =1 v j where dij = length of the ith ray within the jth cell. In a survey which incorporates m travel times, acquired for rays at various positions and orientations across the region, there will be m such summations. These can be expressed in matrix form as: t = Dw (4.12) where D is an m × n array having elements of the form dij; w is an n-element column vector containing the current estimate of the reciprocal velocities (‘slownesses’); and t is a column vector of the m travel times calculated across the discretized velocity field. Let p be the column vector of the m observed travel times from a field survey. For reasons to be discussed in detail later, it is usually not possible to find a velocity field which is exactly consistent with the measured travel times. Thus the process of reconstructing a velocity field is equivalent to the minimization of a residual vector, e, defined by the equation: (4.13) e = p − Dw Commonly used strategies for computerized geotomographic reconstruction include the damped least squares method (Bois et at. 1971), the Back Projection Technique, BPT (Kuhl and Edwards 1963; Neumann-Denzau 1984), the Algebraic Reconstruction Technique, ART (Kaczmarz 1934; Peterson et al. 1985), and the Simultaneous Iterative Reconstruction Technique, SIRT (Dines and Lytle 1979; Gilbert 1972), although many others are available (for example, Gordon and Herman (1974)). The acquisition diagram in Fig. 4.25 shows straight rays. In practice, rays may deviate from straight paths as waves undergo, for example, refraction in an heterogeneous velocity field. Therefore the path followed by a ray, between a source and receiver, is not known. D is a function of w: the elements of w and D are unknown. Thus reconstruction involves not only the distribution of errors along ray paths, but also the determination of an appropriate position for that ray path. The product of reconstruction is a single velocity for each cell. These velocities may be contoured and displayed as colour or grey scales in a tomogram. A seismic velocity tomogram is, necessarily, an approximation. It is a two-dimensional, discrete estimate of a continuously varying, three-dimensional function — that is, the distribution of the seismic velocity properties of the ground. The accuracy of a tomographic estimate of the subsurface and, hence, its usefulness as a predictive tool is influenced by many diverse factors. These affect the ability of the technique to determine the size, form and seismic velocity (and hence stiffness) of features in the ground.

Idealizations of wave behaviour Seismic energy travels through the ground as waves. There are a number of different physical theories associated with the description of wave behaviour. Each particular theory or idealization may be useful in one context, but it is often necessary to appeal to more than one theory to understand fully how seismic energy can be transmitted through the ground. For example, a source of seismic energy at a point in a homogeneous isotropic medium will produce a set of spherical wavefronts and these can, as a simplification, be represented as a set of rays emanating from the source (Fig. 4.26a). The ray approximation is convenient for

45

Site Investigation use in geotomographic reconstruction because of the line integral relation between propagation velocity and travel time assumed in eqn 4.11. According to Huygen (who first put forward a wave theory for light), each point on a wavefront can be regarded as a possible secondary source. This concept leads directly to Fermat’s principle of stationary travel times, which identifies the ray path as that giving the minimum travel time. A corollary of Fermat’s principle is Snell’s law of refraction, which governs the deviation of a ray at a velocity interface. Snell’s law is restrictive because it only permits energy propagation normal to a wavefront. Ray paths for reconstruction determined using Fermat’s principle will allow for reflection, refraction, diffraction, and also head waves.

Fig. 4.25 Definition of notation for reconstruction. Figure 4.25, with its straight ray paths, gives an oversimplified picture of energy passing from seismic source to receiver. In situations where there is not much variation in seismic velocity in the region of interest, the straight ray assumption may be reasonable. ISRM (1988) suggests that this is the case when velocity contrasts are less than 20%. When there are greater velocity contrasts, straight-ray reconstruction may lead to unacceptably inaccurate tomograms. Wave energy may be refracted (Fig. 4.26b) or diffracted (Fig. 4.26c) around obstacles, and head waves may form along velocity interfaces (Fig. 4.26d). Attempts to reconstruct the distribution of seismic velocities will be doomed to failure if the mathematical model being used assumes that rays can only bend according to Snell’s Law, but diffraction or head waves dominate. To illustrate these points, consider the two tomographic surveys outlined below. The first was designed to locate a tunnel crossing the tomographic plane. If ray paths are assumed to be those permitted by Snellian refraction then there should be receivers which are unable to detect any energy emitted by the source (Fig. 4.27a). In practice, all receivers detected incoming wave energy, albeit at various reduced amplitudes. The presence of the water table produced a higher velocity layer a short distance below the tunnel and gave rise to head waves which provided a route around the void (Fig. 4.27b). Even without diffraction and head waves, if refraction is invoked then many rays will travel around the anomaly (Fig. 4.27c). As might be expected from the discussion above, a P-wave survey failed to show any signs of the tunnel (Fig. 4.27d).

46

Subsurface Exploration: Engineering Geophysics

Fig. 4.26 Wave propagation.

Fig. 4.27 Unsuccessful tunnel location.

47

Site Investigation

An idealization of tunnel geometry is shown in Fig. 4.28. Simple mathematics can demonstrate that for any given size ratio (a/L) there is a critical velocity ratio (V2/V1) below which an increase in velocity contrast has no effect on travel time. At lower velocity ratios the velocity (and hence stiffness) of the inclusion (V2) cannot be recovered correctly during reconstruction, because first arrival energy does not travel through the lower velocity material. For the simple geometry shown, it is obvious that the maximum effect that a lowvelocity inclusion may have is to double the travel time. When the inclusion occupies as much as 30% of the distance between the source and receiver, the travel time will increase by no more than 6% of the value when no low-velocity inclusion is present, regardless of the velocity contrast. The detectability of high-velocity inclusions will not he influenced by this effect.

Fig. 4.28 Normalized travel time ratio as a function of velocity contrast. As a second example of problematic wave phenomena, Fig. 4.29 shows two adjacent shearwave tomograms that were obtained from the London clay. The diagonal features which appear in these images result from picking first-arrival shear-wave energy arising from tube waves propagating in the water-filled source boreholes. It is emphasized that these are not a feature of the ground. Virtually identical features can be synthesized by assuming that a compressive wave propagates down the hole in the borehole fluid, generating a shear wave upon its impact with the base.

48

Subsurface Exploration: Engineering Geophysics

Fig. 4.29 Atefacts produced by tube-wave propagation. Finally, there are other aspects of wave behaviour that are important. For example, most tomographic reconstruction is based on the assumption of plane structures in the ground. It should always be remembered that seismic waves are three-dimensional in form and that first arriving seismic events may have followed paths outside the plane of the boreholes. In addition, features in the ground which are smaller in dimension than about one wavelength of the signal will not be capable of being resolved by tomographic reconstruction. In our experience, P waves typically will be expected to have wavelengths of between 1 and 5 m, while for S waves the wavelength is of the order of 1 to 3m.

Influence of data errors Data errors can arise because of inaccurate seismograph triggering, as a result of mis-picking travel times, and also from locating the seismic source and receiver stations incorrectly. The latter class of error will usually be negligible in cases where a borehole deviation survey is available. Any estimate of travel time between source and receiver will involve some level of error. This error is a function of the accuracy with which the source trigger time is known, and. the certainty with which first arrival events can be identified on the seismic records. The latter is largely a function of signal-to-noise ratio. The influence of observational errors in the travel time data set, p, on the spatial resolution of a survey can be quantified as follows. Let terr be the empirically determined uncertainty on these data. An inclusion would be imperceptible if the increase in travel time due to its presence were less than the travel time error. Thus the following condition must be satisfied: t err < Rt (4.14) t where t is the observed travel time for the wave and Rt is the ratio of travel times for the feature, for example, as defined in Fig. 4.28. Rt is a function of the geometry of the problem and the velocity ratio; values for a simple square inclusion can be obtained from Fig. 4.28. If eqn 4.14 is not satisfied, data errors will effectively mask the presence of the feature. In practice, estimates of terr can be based on previous field experience, allowing for signal degradation caused by ambient noise levels and attenuation due to increased borehole spacing. For example, in a survey across a 15 m span of London clay, the authors have observed travel time and errors of the order of 10 ±0.2ms (terr/t=2%) and 75 ±0.8ms (terr/t=1%) for P and S waves, respectively. These values suggest that, to be detectable, a void must have dimensions greater than about 13—18% of the borehole spacing.

49

Site Investigation As has been indicated, in a typical geotomographic survey, the matrix D does not have full rank. Furthermore, D is ill-conditioned. Ill-conditioned systems tend to magnify the effects of data errors (Jackson 1972). This problem can be reduced by identifying and excluding outliers from the data set and also by ‘damping out’ the influence of the smaller eigenvalues in the tomographic system. Nevertheless, to some degree, observational errors will, unavoidably, affect a tomogram that is derived from field data. An effective empirical method of assessing the influence of travel time errors on a reconstructed image is given by the following procedure. A numerical model of the suspected velocity field is simulated. Using a suitable ray-tracing algorithm, a set of travel times for theoretical rays across the field are calculated. A second data set is generated by adding random ‘error’ values, limited in magnitude by terr, to the travel times in the first set. Tomograms are reconstructed from both data sets. Subtracting, cell by cell, the reconstructed velocity values in each image will result in a ‘difference’ tomogram. This image indicates the effect of data errors in a particular tomographic system, as processed by the chosen reconstruction algorithm. The image can be interpreted thus: if cell velocities in the difference tomogram show a deviation of, say, 5% from the velocities within adjacent cells, then such fluctuations in a field tomogram should, perhaps, be attributed to observational error rather than genuine variations in the properties of the surveyed region. As with many new and complex techniques, there is a danger that tomography will become discredited as a result of thoughtless misuse. As we have demonstrated above, there are potentially many reasons why the technique might be expected not to succeed, given particular site conditions. If tomography is to be successful then a number of basic criteria should be applied during the planning and design of the work. The preliminary design of a tomographic survey should consider: 1. the type of wave to be used (i.e. P or S). This decision should be made on the basis of the expected velocity contrasts in each case, and the predicted wavelength in relation to any target; 2. the expected signal-to-noise ratio, and the repeatability with which the seismograph can be triggered by the source; these will influence the travel time errors; and 3. the size and geometry of the required tomogram, based on the amount of ground to be investigated, and the expected size of the target. It is essential that shallow tomographic surveys are planned on the basis of a reasonable knowledge of the likely range of ground stiffnesses (and hence seismic velocities), and the information required from the survey. If tomography is intended to detect a ‘target’ (for example, a fault or a cavity) then the possible orientations and sizes of the target should be estimated. Following this, synthetic travel times should be generated for a number of possible survey geometries (i.e. borehole separation, and the down-hole source and receiver spacings) and processed using a range of reconstruction techniques. In this way, the design of the survey may be optimized and the viability of the technique assessed. On the basis of our experience to date, the following may be stated. 1. Low-velocity anomalies (such as cavities) appear more difficult to detect than highvelocity anomalies (for example, hard inclusions). 2. The absolute values of seismic velocity recovered from a survey may be regarded as indicative of ground stiffness variations, but should not be used in an absolute way in engineering calculations. 3. Velocity variations are likely to be most reliably reconstructed when velocity contrasts are low.

50

Subsurface Exploration: Engineering Geophysics 4. Planar, or approximately planar features (such as faults), can relatively successfully be located, provided that they strike normal to the tomographic plane. An example of the successful reconstruction of real data, over a fault in an oil storage cavern floor, is shown in Fig. 4.30.

Fig. 4.30 Successful application of geotomography: detection of a fault in weak rock.

Seismic tomographic surveys Tmographic surveys are commonly carried out between two boreholes. In the simplest case an energy source is placed in one borehole (transmitter borehole) and an array of receivers placed in the second borehole (receiver borehole) as shown in Fig. 4.31. The survey is performed by keeping the energy source stationary in the transmitter borehole and moving the receiver array up or down the receiver borehole taking a record for each receiver array position. The energy source is then moved to a new position and the process of moving the receiver array is repeated. When the energy source has traversed the section of interest the survey is complete. In cases where the energy source used may give rise to errors in travel time determination because of triggering problems the survey may be carried out using two receiver boreholes. In some cases only one borehole is used. Here the energy source is placed on the surface at different distances from the borehole and the receiver array is run in the borehole. This arrangement has proved very useful on restricted sites or in cases where tomography was not considered in the original design of the site investigation. The tomogram produced by this arrangement has a characteristic triangular shape. Figure 4.30 is an example of a tomogram produced using this arrangement.

Borehole preparation When carrying out seismic tomographic surveys with either P waves or S waves the boreholes must be lined with plastic (ABS) casing. The use of plastic casing is critical to the success of the survey because metal casing presents significant velocity contrast with the surrounding ground. The casing is capped at the base and inserted into a grout-filled borehole. The grout is

51

Site Investigation necessary to provide acoustic coupling between the borehole and the ground. A bentonite or cement-bentonite grout is generally suitable.

Fig. 4.31 Acquisition geometry for a typical seismic tomographic survey. In order to minimize errors in tomographic reconstruction it is necessary to define with reasonable accuracy the three-dimensional geometry of the survey section. In order to achieve this borehole verticality surveys must be carried out as part of the tomographic survey.

Equipment and field techniques The basic equipment required for most tomographic surveys includes:

• • • • •

energy source (P wave or S wave); receiver array; compressed air for use with clamped sources and receivers; winches for lowering tools down the boreholes; and seismograph.

The equipment in terms of energy source and receivers will depend to a large extent upon the type of seismic wave being employed for the survey. Table 4.8 outlines the equipment commonly used for P- and S-wave surveys. The borehole sparker operates in a similar manner to that used for continuous seismic profiling over water (see Table 4.7). An electrical discharge of some 4kV is generated at a known depth in the transmitter borehole. With such high voltages being used it is important that site personnel are kept clear of the high tension cables crossing the site, and that only suitably trained personnel are permitted to operate the equipment. The signal produced by the sparker has a sharp leading edge and is highly repeatable which aids picking travel times. Since the energy source is electrically operated, an electrical signal can be sent to trigger the seismograph the same instant as the electrical discharge occurs in the borehole. This means

52

Subsurface Exploration: Engineering Geophysics that timing errors associated with triggering are almost eliminated and surveys may be carried out using only two boreholes.

Table 4.8 Equipment commonly used for P- and S-wave cross-hole tomographic surveys Number of Comments Type of Energy source Type of receiver boreholes seismic required wave P Borehole Hydrophone 2 Borehole sparker is a very sparker (freely repeatable energy source with (freely suspended) negligible trigger timing errors. suspended) S Clamped Clamped three- 3 For most commercially shear wave component available mechanically operated hammer geophones shear wave hammers two receiver boreholes are necessary for accurate travel time measurements. When carrying out P-wave surveys using a borehole sparker it is common practice to use an array of hydrophones in a water-filled receiver borehole. Because shear waves cannot travel through water, only P waves will be received at the hydrophones. The number of hydrophones in the array will depend upon the number of channels available on the seismograph. Typically arrays will comprise 10 or 20 hydrophones. The spacing between hydrophones may be changed easily on site, and because the array is freely suspended within the borehole moving it from one location to the next may be done very rapidly. Shear waves are generally generated using a mechanically operated hammer striking an anvil which is clamped to the borehole wall. A typical hammer used for generating vertically polarized shear waves in boreholes is the Bison hammer shown in Fig. 4.32. Coupling between the hammer and the ground is provided by clamping the anvil to the borehole casing. The clamps are operated pneumatically and retracted using springs. As a result of the mechanical operation of the hammer, the seismograph must be triggered by placing a piezoelectric transducer or a hydrophone in the transmitter borehole close to the hammer. The triggering system often produces significant timing errors. In order to avoid such timing errors it is common practice to use two receiver boreholes such that travel times are measured between them rather than between transmitter and receiver. The receiver array for shear wave surveys comprises a series of three-component geophones. A three-component geophone consists of three geophones orientated in three mutually perpendicular directions (one vertical and two horizontal). The operation of geophones is described later. A typical receiver array includes four three-component geophones each of which is clamped independently to the casing using the same system as for the hammer. In order to ensure that the alignment of each set of geophones remains fixed within the borehole the array is contained within a flexible (normally rubber) tube. The tube is lowered down the borehole using a set of rods so that the alignment may be controlled such that one set faces in the optimum direction to maximize the amplitude of the incoming signal. This arrangement makes it impossible to change the spacing of the receivers on site. The borehole spacing and measurement interval used in tomographic surveys are similar to those used in conventional cross-hole seismic surveys which are discussed later. The time taken to conduct a tomographic survey will depend upon: 1. the length of the section being surveyed; 2. the station interval;

53

Site Investigation 3. the number of receivers in the array; 4. whether the source and receiver array are clamped or freely suspended; and 5. the amount of stacking required to reduce the signal-to-noise ratio to an acceptable level. A 40 m section may be surveyed in a day using a borehole sparker and ten hydrophones with a station interval (i.e. transmitter interval and hydrophone spacing) of 2 m and between 4 and 8 stacks per record. The same section may take at least two days using a clamped source and receiver. The field work required for even a minor tomographic survey is more than that required for most conventional cross-hole or surface seismic surveys. However, the major cost in tomographic surveys is in the data handling and processing. Generally several hundred travel times must be picked for each section surveyed. These must be checked and entered into the computer for use by ray tracing algorithms.

Fig. 4.32 Bison shear wave hammer.

54

Subsurface Exploration: Engineering Geophysics

DETERMINATION OF PROPERTIES

Geophysics is generally of very little use in providing parameters for geotechnical analysis and design. There is, however, one important exception. As already noted above, several seismic methods can be used to obtain information on the stiffness of the ground. Traditionally, the seismic methods preferred by geophysicists have been based upon methods used for deep mineral exploration, and they have therefore relied upon P waves. Compressional (or primary) waves are convenient for these purposes because they can be readily generated (for example using hammer blows or explosives), and their identification on the seismic record in simple — they are the first arrivals. Deep rocks have high skeletal stiffnesses, and therefore the P-wave velocity of the rock is usually much greater than that of its interstitial fluid. In geotechnics, however, the materials to be tested are generally of low effective stiffness (i.e. their skeletal stiffness is low), and are often fissured and fractured. Furthermore, in most temperate countries the groundwater table is close to ground level. Therefore the P-wave velocity of near surface soils and rocks is normally close to that of water (about 1500 m/s), and P-wave velocities cannot be used to distinguish between the different types of ground or to provide data on the stiffness of the ground (rather than the stiffness of its pore water). Recent work has therefore concentrated upon the use of seismic shear waves to characterize soils and weak rocks. Shear waves are more difficult to detect, since they occur as secondary events, arriving some time after the primary waves. But they have two helpful characteristics — their amplitude is often large, particularly if a special shear-wave source (which is designed to produce energy which is rich in shear waves) is used, and their sense can be reversed. It is standard practice to reverse shear-wave inputs in order to help the identification of first arrivals, since compressional waves do not reverse. As was noted above, the very small strain stiffness of the soil or rock mass is linked to seismic shear-wave velocity by the simple equation: 2 G = Vs ρ (4.10) and because the density of soils and rocks can usually be estimated with reasonable accuracy, the shear modulus (Go) can readily be determined from the shear-wave velocity. Until recently, it was thought that, because of the highly non-linear behaviour of most soils (Fig. 4.33), the very small strain stiffness determined from geophysics was so much greater than the values required for design that it was of no use in geotechnical design. It is now known that in many cases this is not so. Finite element analyses of the deformations around civil engineering structures have repeatedly shown that strain levels at working loads are very small, of the order of 10-2-10-1% strain, and this has led to a realization that stiffnesses obtained at higher strains during conventional laboratory triaxial tests are far too low. At the same time that special laboratory techniques have been introduced there has been a rapid growth in the use of seismic shear-wave geophysics. Comparisons of the shear moduli obtained from geophysics with those obtained from the back-analysis of observed deformations around full-scale structures has shown that while geophysics does overestimate the stiffness of the ground at engineering strain levels, it does so by a relatively low margin. In weak rocks and strongly cemented soils Go may approximately equal the stiffness at engineering strain levels, while in loose granular media and heavily overconsolidated clays it may be between two and three times greater. Given the relatively large uncertainties associated with stiffness determined in other ways, whether in

55

Site Investigation the laboratory or using in situ testing, shear-wave geophysics must be regarded as a very useful site investigation technique.

Fig. 4.33 Stiffness degradation curve for stiff clay from Tokyo Bay (Mukabi et al. 1991). Laboratory measurements of small-strain stiffness have frequently been reported in the literature, and are now commonplace in UK site investigation practice, and there are also well-established more conventional in situ testing techniques which can be used to determine the stiffness of the ground (for example, plate tests and pressuremeter tests), so that it might be thought that there is little point in carrying out further (very small strain) stiffness measurements by geophysics. However, laboratory tests are subject to sample disturbance, and may not be possible in some soils, because sampling is impossible. And both laboratory and in situ tests are generally carried out on small elements of soil, and may not give stiffnesses which are representative of the mass, inclusive of fracturing and the full range of particle sizes present. Below we describe three techniques which are relatively well established, and under normal circumstances (in the absence of severe background noise) can be expected to yield good results.

56

Subsurface Exploration: Engineering Geophysics Seismic refraction The refraction method is based on the critical refraction of seismic waves at the boundaries between materials with different characteristic seismic wave velocities (Fig. 4.34). Snell’s Law governs the refraction of seismic waves across a velocity interface. From Fig. 4.34: sin i Vo (4.15) = sin r V1

Fig. 4.34 Seismic refraction using P-wave events (first arrivals). For critical refraction, sin r = 1 (i.e. r = 900). Hence: 57

Site Investigation Vo (4.16) V1 Critical refraction therefore will only occur if V1> V0, and the rays meet the interface at the critical angle of incidence, ic. If V1 < V0, rays will be refracted towards the normal to the interface making critical refraction impossible. This imposes a limitation on the method and is discussed in more detail later. If a ray meets the interface at an angle less than the critical angle, refraction takes place and the ray continues downwards (Fig. 4.34). Also some of the energy is reflected at the interface. If the angle of incidence exceeds the critical angle, total reflection occurs. sin i c =

The critically refracted ray travels along the interface in the lower (higher velocity) medium. This produces an oscillatory motion parallel to, and immediately below, the interface. Because there is no slippage along the interface, the upper medium is forced to move with the lower medium in a zone adjacent to the interface. This causes the generation of continuous new disturbances which are analogous to sonic booms produced by aircraft flying at supersonic speeds. The shock waves produced are known as head waves. It can be shown that the resultant head wave fronts emerge at the critical angle 4, (Dix 1939). It is the detection of the head wave by the geophones that enables data to be obtained for layers other than the surface layer. The seismic record produced by a geophone placed at a given distance from the energy source will show a combination of direct events,2 refracted events (head waves), reflected events and ‘ground roll’ (surface waves). The reflected waves will always arrive later than the direct waves or the refracted waves. In one special case, the refracted wave and the reflected wave arrive simultaneously. The energy source to geophone distance at which this occurs for a simple two layer case (Fig. 4.34) is given by: (4.17) x = 2Z 0 tan ic where x = distance between energy source and geophone, Z0 = depth to first refractor and ic = critical angle. At distances less than 2Z0 tan ic the head wave does not exist and hence no refracted events appear on the seismic record. In Fig. 4.34 the geophones nearest the energy source (shot point) will record the direct wave as a first arrival event. However, because the critically refracted wave travels at a greater velocity than the direct wave, eventually the head wave produced at the interface will arrive at the geophones before the direct wave. The distance from the shot point at which the direct wave and the head wave arrive at the ground surface simultaneously is termed the critical distance X0. The critical distance is used in the interpretation of seismic refraction data. Other critical distances may be defined for multilayer systems for the simultaneous arrival of refracted waves from different layers. The equipment required for refraction surveying includes an energy source, geophones, a take-out cable, and a signal-enhancement seismograph. The fundamental requirement for an energy source for seismic surveying is that it should be capable of delivering to the ground an impulse which has a sharp leading edge, so that firstarrival events can be accurately picked from the geophysical records. Traditionally seismic refraction has been used primarily to determine sub-soil geometry, and has used compressional wave arrivals. For this purpose, and for investigation of shallow depths, it has been sufficient to use a sledge hammer striking a metal plate placed on the ground or, for greater depths of penetration, detonators placed in shallow holes. Figure 4.35 contrasts this 2

The signal produced by a geophone in response to a wave front is termed an event. 58

Subsurface Exploration: Engineering Geophysics with the techniques required to produce shear waves. For shear waves two sets of data are required at each source position, in order to give reversals (Fig. 4.36). Traces can be stacked, as with P-wave surveys, but only for shear-wave energy input in the same direction. A commonly used borehole shear- wave source is the Bison hammer, shown in Fig. 4.32.

Fig. 4.35 Simple methods of producing shear-wave energy for shallow seismic surveys. The ground motion induced by the passage of the seismic wave is detected by a small electromagnetic transducer, termed a geophone. Typically either 12 or 24 geophones are used. A geophone consists of a coil of wire, suspended between the poles of a magnet which is fixed to its outer casing. The coil acts as an inertia element, such that any vibration causes the coil to oscillate within the magnetic field and generate an output voltage. The output of the geophone is therefore proportional to the velocity of the ground on which it is placed. Geophones can be mounted either horizontally or vertically, and for shear-wave surveys should be orientated in the direction of the incoming shear-wave energy, in order to maximize the signal. The seismograph is connected to the geophones via the take-out cable. Signal enhancement seismographs have the ability to stack repeated geophone traces, and can normally accept up to 12 or 24 geophones. They amplify the small electrical signals that the geophones produce as a result of ground vibrations, and use a precise common time standard against which to record the trace produced by each of them. After amplification the signal is digitized, either using a fixed gain (for example in the ABEM Terraloc) or with ‘automatic gain’ as in Digital Instantaneous Floating Point (DIFP) machines such as the Bison 9000 series seismographs. When using machines with a relatively small dynamic range (the Terraloc Mk II has only eight-bit resolution) it is necessary to set the gain carefully, and to use different records to capture compressional and shear-wave arrivals. DIFP seismographs overcome these problems. Geophones are normally laid out at regular intervals in a line (at points 1 to 12 in Fig. 4.34), termed a ‘spread’, and are orientated in the best direction to receive maximum incoming shear-wave energy. The shot point is normally colinear with the spread. A minimum of two

59

Site Investigation shot points should be used, one at each end of the spread, and at each shot point the direction of energy input must be reversed.

Fig. 4.36 Shear wave reversals on two traces recorded by a signal-enhancement seismograph. In order to determine the seismic velocities of different materials it is necessary to determine the travel time from the shot position to each geophone for the relevant type of wave. When the shot is made, either by hammer blow or explosives, the seismograph is triggered and starts recording for a short predetermined period, It is advisable to place a geophone next to the shot point in order to provide a reference time, as well as a check on triggering accuracy. The traces that are obtained may be viewed on screen or printed, immediately after the shot, in order to guarantee that good quality data have been obtained. They are best processed by computer. For P-wave surveys the first break is picked, normally by eye. For shear-wave surveys two traces, with the same shot point but with the energy input in opposite directions, are superimposed. The traces are shifted with respect to time to obtain a match on the rise of the trace at the reference geophone. The trace is then searched for the point at which reversals

60

Subsurface Exploration: Engineering Geophysics first occur, and this time is then picked. A time-distance graph is plotted for either the P- or Swave arrivals, or for both. For multi-layered ground the velocities and depths of the various layers are interpreted as follows. The time—distance graph is split into straight-line portions, as in Fig. 4.37. The gradients of these lines give the velocities of the various layers. The line passing through the origin is produced by direct waves and the line which when extrapolated intersects the time axis (at time T0) is produced by critical refraction in the second layer. The intersection of these two lines defines the distance from the shot at which the direct and refracted waves arrive at the surface simultaneously. This has been previously defined as the critical distance X0. The point at which the extrapolation of the second line (refracted events) intersects the time axis is the intercept time T0. This has no real meaning with respect to the physical model since it is impossible for head waves to emerge at the shot point. It will be seen from Fig. 4.34 that no head waves emerge at the surface between the shot point A and a point B which is 2Z0 tan ic distant from A. The critical distance and intercept time are used together with the characteristic P-wave velocities for each layer to calculate the perpendicular distance Z0 between the shot point and the top of the second layer (Fig. 4.34). The distance Z0 may be determined using critical distance by: X V − V0 ( 1 ) Z = (4.18) 2 V1 + V0 It is, however, easier to determine the intercept time from the time-distance graph, and hence depth determinations are normally made using this parameter from the expression: V1V0 T (4.19) Z = 2 (V 2 − V 2 ) 1

0

However, when experimental errors which affect the determination of V1 are considered (Steinhart and Meyer 1961), it is preferable to use the critical distance in depth determinations. Figure 4.37 shows a time-distance graph for a multi-layer case, Intercept times and critical distances (simultaneous arrival of two refracted events for layers other than the surface layer) may be determined for each layer. The above equations can be extended to allow the depth of each layer to be calculated. The accuracy of velocity and depth determinations together with the chances of actually detecting different strata (or other geological bodies) are very much dependent on velocity contrast between different media. Providing that the velocities of the layers increase with depth, in general the greater the velocity contrast the greater the confidence in identifying different strata and the greater the accuracy of depth determinations. Clearly it is important to have at least some idea of the expected materials before making the choice of using the refraction method, as this method may not be suitable. If the velocity contrast is low (e.g. V1: Vo 3:1) the head wave inclination (with respect to the normal to the interface) is very small. This causes the velocity term in the depth determination (i.e. V1Vo/ (V12-Vo2)) to be a minimum, and hence accuracy is increased by minimizing errors due to erroneous velocity determinations.

61

Site Investigation Bullock (1978) states that the accuracy of depth determinations is expected to be within ±15% of the true depth over a range of depths of interest between 3m and 100 m.

Fig. 4.37 Time—distance graph for a simple multi-layer case. The interpretation of the time—distance graphs shown in Figs 4.34 and 4.37 is straightforward. In practice, however, the interpretation is often more complicated. The features which commonly give rise to a complicated time—distance graph include: 1. dipping interface; 2. irregular interface and buried channels; 3. lateral changes in velocity;

62

Subsurface Exploration: Engineering Geophysics 4. buried step faults; and 5. lateral changes in stratum thickness. These features may be identified if sufficient data are obtained from each geophone spread. This involves using more than one shot point for each spread in order to maximize the amount of data. It will not normally be worthwhile, during geotechnical investigations, to carry out the additional work necessary for this. In many near-surface situations the shear-wave velocity of the ground increases approximately linearly with depth, as weathering of the ground reduces. Seismic refraction techniques can then be used to get reasonable approximations to the very small strain stiffness of the soil or rock beneath the surface. Whilst, as we have seen, the customary method of interpreting seismic refraction data is to assume that the ground is layered, each layer having a constant velocity, an alternative treatment is to fit the data to an inverse sinh function, corresponding to a linear increase of velocity with depth (Dobrin 1960). In this case the refraction paths are arcs of circles (Fig. 4.38) and from the geometry it can be shown that the travel time T is: 2 kx (4.20) T = sinh −1 ( ) k 2Vo where V0 = velocity at the surface, x = horizontal distance from the shot point to the geophone, k = increase in velocity with depth, and V = velocity at any depth, given by: (4.21) V = Vo + kz Abbiss (1979) fitted the above equations to data obtained from P-wave seismic refraction surveys on the fractured chalk at Mundford by Grainger et al. (1973), and using the relationship:

V z= o k

kx 1+ 2Vo

1/ 2

−1

(4.22)

where z = depth reached by the survey and x = distance from the shot point to the geophone, obtained good correspondence between the velocities from this method and those from the layer method. Abbiss calculated the dynamic moduli obtained from this method of interpretation, using: 2 (4.23) E d = G 2(1 + ν ) = V s ρ 2(1 + ν ) Given that: Vs (1 − 2ν ) = (4.24) V p 2(1 − ν ) 2

Ed = V p ρ

(1 − 2ν )(1 + ν ) (1 − ν )

As this demonstrates, if compressional (P) waves are used, then Poisson’s ratio must be known. Although this is not normally the case, in this instance there had been both laboratory and field measurements which gave a value of Poisson’s ratio (v) of 0.24 (Burland and Lord 1970; Burland et at. 1973). The dynamic moduli obtained from the seismic method were about twice those backfigured from the movements below a large instrumented tank test.

63

Site Investigation

Fig. 4.38 Seismic ray paths for a linear increase of velocity with depth (Abbiss 1979). Cross-hole and down-hole seismic surveys The down-hole seismic method, used in conjunction with the in situ cone test, has been described under ‘Profiling’, earlier in this chapter. Cross-hole surveys are described in detail in Ballard et al. (1983). They are typically carried out using three parallel in-line boreholes, with plastic lining (typically about 100mm internal diameter) grouted into them. Horizontal borehole spacing is typically 5—7 m. The closest spacing possible is desirable, in order to achieve a high signal-to-noise ratio, but this consideration must be tempered by the need to record accurate travel times between the boreholes. A vertically-polarized shear-wave source, such as the Bison hammer, is lowered into one of the end holes. Seismic waves are generated by raising or dropping the frame of the hammer against the central, clamped shuttle. In the standard hammer, wave initiation is detected by a piezoelectric transducer, and this is used to trigger a seismograph. Three-component geophones are clamped at the same level as the hammer in the other two boreholes. The seismograph records the incoming waves at the two holes, and the travel time between them is obtained by subtraction. All three boreholes are surveyed for verticality, so that their relative positions are known precisely at each test depth. The shear and compressional wave velocities are then calculated by dividing the appropriate travel time by the distance between the receiver boreholes at that depth. Measurements are made at depth intervals of between 1 m and 5m, depending upon the total depth to be surveyed. It is better to take measurements at intervals of 1 m even in deep boreholes, because the data so-obtained can then be averaged using a rolling average method, so reducing the impact of any ‘rogue’ measurements. As is usual when picking on first breaks, about 10—15 traces are stacked during data acquisition, to improve the signal-to-noise ratio. Boreholes constructed, for cross-hole surveys can also be used for up-hole and down-hole surveys. In 911e case of down-hole surveys the speed of surveying can be greatly increased by using strings of three-component geophones. Typically three or four geophones are used in each string, with three orthogonally orientated geophones at each level. The sources of P and S waves are deployed at the surface, and typically consist of vertical hammer blows on a metal plate (for P waves) and horizontal hammer blows on a loaded plank (for S waves). As with the cross-hole survey, the interval velocity is determined from the difference in travel time and the distance between pairs of geophones, but in this case the source is located within a metre or so of the top of the borehole, and the time intervals are determined between adjacent geophones in a single borehole. In both types of survey geophones should be orientated in the borehole so that one set faces in the optimum direction to maximize the amplitude of the incoming signal. Figure 4.39 shows the results of cross-hole and down-hole seismic surveys in Mercia Mudstone. The material at this site is strongly layered, containing evaporite materials in the form of veins or bands of gypsum, and layers and scattered nodules of gypsum, dolerite and anhydrite. It can be seen that the cross-hole velocities are consistently higher than those from

64

Subsurface Exploration: Engineering Geophysics the down-hole survey, where the geophones were at 6.4 m centres. It might be thought that this was related solely to anisotropy, but in layered ground such as this it is a product, at least partly, of the method of test. Both methods use first arrival events in order to determine the velocity of the ground. But as Fig. 4.40 shows, the cross-hole travel times are, on average, faster because the seismic energy being recorded as first breaks travels through the faster layers, in the form of head waves. If the purpose of the surveying is to determine the average ground stiffness, via the seismic velocities, then the down-hole velocities will provide a more reliable estimate than the cross-hole values in horizontally layered ground, since they average the properties of the materials between the geophones. On the other hand, cross-hole velocities are more suited to identifying layering in the ground. This makes it clear that crosshole and down-hole seismic results should only be interpreted in conjunction with good borehole records, and that they should be regarded as complementary methods.

Fig. 4.39 Seismic velocities (Vs and Vp) for Mercia Mudstone, determined from cross-hole and down-hole measurements (Pinches and Thompson 1990).

65

Site Investigation

Fig. 4.40 Biassing towards higher seismic velocities in cross-hole seismic tests, as a result of head waves (Pinches and Thompson 1990). The surface wave technique As noted at the start of this chapter, seismic energy travels through the ground as both body waves (longitudinal compressional (P) and shear (S) waves) and as surface waves (Love waves and Rayleigh waves). Energy sources used in seismic work are not generally rich in Love waves, but near-surface measurements can be affected by Rayleigh waves, which travel at a similar (although slightly slower) velocity to shear waves. Rayleigh waves are associated with a particle motion which is elliptical in the vertical plane, and which attenuates rapidly with depth and with distance from the source. A frequency controlled vibrator, which may be relatively lightweight ( σh):

∆σ 1 = −σ v and K 0 =

(σ h − u 0 ) (σ v − u 0 )

∆σ 3 = −σ h

(6.3) (6.4)

For a saturated clay B=1, therefore: ∆u = u k − u 0 = ∆σ 3 + A( ∆σ 1 − ∆σ 3 )

(6.5)

Under elastic soil conditions it can be shown that A =⅓ , and therefore the above equation can be rewritten in terms of the pore pressure expected from an ‘elastic clay’ and the difference of a real soil from this value, i.e.

1 1 ∆u = [(∆σ 1 + ∆σ 3 ) + ( A − )(∆σ 1 − ∆σ 3 )] 3 3 ∆σ 1 − ∆σ 3 = − p ′K 0 therefore:

1 ⎤ ⎡ (1 + 2 K 0 ) ∆u = − p ′⎢ + u 0 ⎥ − p ′( A − )(1 − K 0 ) 3 3 ⎦ ⎣

(6.6)

8

Site Investigation

Now: ⎤ ⎡ (1 + 2 K 0 ) ⎛ 1⎞ − p k′ = −u k = −( ∆u + u 0 ) = p ′⎢ + ⎜ A − ⎟(1 − K 0 ) ⎥ 3 3⎠ ⎝ ⎦ ⎣

(6.7)

If, as is approximately the case for heavily overconsolidated clays, the material behaves elastically during unloading:

uk = − p′

(1 + 2 K 0 ) 3

(6.8)

Skempton and Sowa (1963) carried out experiments on Weald clay to find the differences between predicted and observed effective stress levels after stress relief under laboratory conditions. The resulting values are shown in Table 6.8, where p ′ is the average effective stress on the soil before stress relief. Thus for this case uk equalled 0.6 p ′ to 0.7 p ′ . Table 6.8 Differences between predicted and observed results for Weald clay Elastic prediction Remoulded clay results p k′ / p ′ 0.73 0.58 p k′ / p ′ 1.00 0.80

Compaction

In granular soil, permeability is high, and therefore vibrations and compressive forces applied to the soil, whether in the ground or in the sampling tube, can lead to changes in density. These effects are most severe in loose granular material, where density will be increased. Compaction leads to changes in the effective strength and stiffness parameters of the soil.

Soil disturbance during drilling

Swelling can occur at the base of the borehole before insertion of a sampler tube, during the taking of a sample, and after sampling when the soil is inside the sampler tube. As examples, the ingress of water to material in the base of a borehole in London clay makes the recovery of soil using a claycutter more difficult, presumably because of the loss of shear strength as a result of swelling; in contrast, a waiting period after sample driving is sometimes used to improve the chances of recovery of London clay in an open-drive sampler with inside clearance. In the second case, swelling increases the diameter of the clay core inside the tube while increasing the effective stress level at the clay/cutting shoe contact. The amount of swelling that can occur is proportional to the change of total stress occurring at the base of a borehole. Thus if the borehole is substantially empty of water there is likely to be more swelling than if the borehole is kept full of mud or water. Total vertical stress changes can effectively be halved by keeping boreholes full of water. The higher the water-table and the softer the soil, the greater is the benefit of a water filled borehole. Figure 6.2 shows the results of analyses (assuming elastic soil with K0 equal to 1) to calculate the variation of pore pressure change caused by borehole stress relief with depth below the base of the hole. It can be seen that large negative pore pressures will be induced, and that these will vary with depth. The vertical extent of pore pressure decrease (and therefore swelling) will be about one borehole diameter.

9

Sampling and Sample Disturbance

The factors which complicate the control of swelling are time and water-table position. If drilling and sampling take place quickly, then little time will be available for water to penetrate the soil. Swelling will be limited. Above the water-table there may be relatively little water available in the borehole, and swelling may be slowed down. The recommendations of Hvorslev (1949) and Idel et al. (1969) with regard to the use of fluid filled boreholes are given in Table 6.9. Table 6.9 Recommendations for the use of fluid filled boreholes Hvorslev (1949) Idel et al. (1969) Boring above groundwater level Keep borehole dry, or use Use water balance. drilling fluid (mud). Water is not permitted. Use water balance. Boring below groundwater level Fill borehole with water or mud, at least when in soft or cohesionless soils. In stiff soils, borehole may be kept dry, but it must then be completely dry.

Hvorslev (1949) commented that ‘swelling … may require considerable time for full development’. Ward (1967) however stated, with regard to the taking of block samples from the Ashford Common Shaft: I was never convinced that we ever appreciated what the London clay was like before the shaft was dug. We showed, even in the short time required to cut out samples, that the blocks which were integral with the base of the excavation were statistically wetter than the pieces which we trimmed off the surface of the excavation while preparing the block.

Fig. 6.2 Stress changes below the bottom of a borehole (modified from Galle and Wilhoit (1962) by Hopper (1992)).

10

Site Investigation If, as is current practice in the UK, water or mud balance is not used in stiff over- consolidated clays, then drilling must occur quickly immediately above each proposed sample location and sampling must take place as soon as that drilling is completed. Compaction, remoulding and displacement of soil beneath or around casing or sampler tubes driven ahead of an open borehole can be minimized if care is taken. Soil displacement can occur as a deliberate method of advancing a borehole; many well-boring rigs operate on the percussion drilling principle, where a heavy drilling bit (referred to as a churn bit) is alternatively raised and dropped by a ‘spudding’ mechanism. This type of displacement drilling leads to significant remoulding and compression of the soil around and ahead of the bit. The depth affected can be up to three times the hole diameter. Similar effects can be unwittingly caused during the more common types of site investigation drilling, principally when using augers or light percussion drilling in soft soil. Most rigs using continuous flight augers are capable of providing considerable downwards thrust; the Acker AD II can give up to about l6000lb (7.2t) while the Mobile B53 can give 19000 lb (8.5t). Over-eager drilling can lead to displacement of soil ahead of the auger before the flights have a chance to remove the soil. In very soft clays the soil may block flights and fail to travel up to the ground surface. Soil displacement then becomes inevitable. Light percussion boring can induce the same sort of problems if casing is advanced below the bottom of the open hole. A plug of soil will form inside the base of the casing and lead to compaction, compression and bearing capacity failure immediately below the bottom of the casing (Fig. 6.3). Casing should never be allowed to go below the bottom of the borehole at any time during drilling; in this case samples taken through the bottom of the casing will probably be highly remoulded if clays, or compacted if sands or gravels.

Fig. 6.3 Displacement of soil beneath casing or a sampler tube (largely after Hvorslev 1949).

11

Sampling and Sample Disturbance The problems discussed above occur equally during the taking of samples. Soil must be displaced to allow the penetration of the sampler tube, and if sufficient shear force is generated between the inside of the sample tube and the soil entering it then the sample may ‘jam’ in the tube. Base heave, piping and caving are all severe effects of stress relief. Base heave can be thought of as foundation failure under decreased vertical stress, and the effects are broadly the reverse of those produced by displacement drilling. When the total stress relief at the base of a borehole is very great compared with its undrained shear strength, plastic flow of soil may take place upwards into the borehole. This effect may be encouraged when pulling sampler tubes out of the soil at the bottom of a borehole. Once flow of soil occurs into the base of a borehole, disturbance may then take place for depths in excess of three borehole diameters ahead of the bottom of the hole and its casing, the actual depth being dependent on the volume of soil allowed to enter the hole. Since base heave is a problem in very soft soils, where the water-table will normally be high, the use of either water or mud balance is recommended. The problem of base and wall instability of boreholes is similar to that of undrained bearing capacity of foundations and the base heave of excavations, which have been the result of considerable research (for example Skempton 1951; Bjerrum and Eide 1956; Britto and Kusakabe 1984). On the basis of this work, Hight and Burland (1990) have concluded that: 1. for an unsupported hole there will be failure of the borehole wall if the undrained shear strength (cu) is less than γD/10, where D is the depth of the hole and y is the average bulk density of the soil above the base of the hole; 2. mud support is helpful in all situations, whereas casing must be continuous, from the top to the bottom of the hole, to be effective; 3. base failure is inevitable in normally consolidated clays; and 4. in lightly overconsolidated clays the factor of safety against base failure will be so low that significant strains will be imposed on the soil immediately beneath the bottom of the borehole. ‘Piping’ is a term used to describe the behaviour of granular soil when its effective confining pressures, and hence strength, are removed as a result of high upward seepage pressures. Under these conditions the individual soil particles are free to move and finer soil particles are carried upwards with the water. The material appears to ‘boil’. When a borehole is inducing total stress relief, and water balance is insufficient to prevent high seepage pressure gradients in the soil at the base of the hole, large volumes of fine granular soil may move up into the casing. Soil below the bottom of the casing will be brought to a very loose state. Piping often occurs when a ‘shell’ is used without water balance, in conjunction with light percussion drilling. It is particularly troublesome if the soil is already loose, the groundwater table high, and the borehole diameter large. The effects of piping on the quality of soil samples taken from granular soil will not normally be too large while light percussion drilling, because loss of fines would be expected when using a shell. Thus, although bulk and jar samples taken from this type of borehole would normally be considered to be quality class 4 (Rowe 1972) in reality they will often be ‘Nonrepresentative’ (Hvorslev 1949). The most serious effects of piping occur because it is normal to use in situ tests to determine the design parameters, such as allowable bearing pressure, for a granular soil. In a borehole, the most common test would be the ‘Standard Penetration Test’ (see Chapter 9), where a 50mm dia. tube is driven into the soil at the bottom of the borehole by repeated blows of a standard weight falling through a fixed distance. The number of blows necessary to drive the tube approximately 300mm is known as the SPT N value, and is used empirically to obtain various soil properties. Piping reduces the density of the soil at the base of the hole, and can therefore give completely false N values; for example, N values have been observed to decrease from 25 blows/300mm to 8 blows/300 mm in sand,

12

Site Investigation which might lead to an unnecessary reduction in allowable bearing pressure for footings from about 250 kN/m2 to 80kN/m2. Sutherland (1963) observed the results shown in Fig. 6.4, where piping appears to have reduced N values by a factor of three or four.

Fig. 6.4 Effects of boiling on SPT ‘N’ values in fine to medium sand (Sutherland 1963).

Piping can be prevented by giving some thought to its causes. The shell or bailer so often used to make progress in granular soils when drilling with light percussion rigs acts by creating suction on the bottom of the borehole. If the shell is a tight fit in the casing then suction will be large, progress will be fast, and disturbance will be enormous. The International Society for Soil Mechanics and Foundation Engineering has prepared a standard for penetration testing in Europe (1977) which specifies the use of a shell with a maximum diameter not greater than 90% of the inside diameter of the borehole casing. This will considerably reduce suction at the base of the hole, but it will not prevent piping if the natural groundwater level is high. When the soil is loose and the groundwater table is high, the borehole should be kept full of water in order to ensure that seepage in the soil at the base of the hole occurs in a downward direction. Under this condition piping cannot occur, provided artesian groundwater is not present. When artesian conditions occur, casing will have to be extended above ground level and drilling may have to take place from a raised platform if piping is to be prevented. If piping is not prevented then the depth of soil affected is a function of the casing or borehole diameter. Fletcher (1965) has discussed the development of the SPT, which was originally used by Colonel Charles R. Gow to provide information on the density of soil formations for the purpose of correlation with experience of bored and driven pile design and installation. In the USA, this test initially used a 52mm dia. SPT tool on size ‘A’ drill rods, in either a 64mm or 102mm casing. The hole was advanced by washboring. British practice currently adopts a minimum hole size of 152 mm; most commonly 204 mm internal diameter casing is used when drilling near to the ground surface in loose deposits. Clearly, British practice is most undesirable because the entire SPT test section can be loosened either by piping, or by stress relief. British SPT N values should be expected, on average, to be lower than values obtained by the American method. Caving typically occurs when boreholes are advanced into soft, loose or fissured soils. Material from the sides of the borehole collapses into the bottom of the hole and must be cleaned out before sampling can take place. Progress is slowed because more material must be removed from the borehole. Stabilization of the sides of boreholes is essential in soils which may collapse or slough. It may be carried out by a variety of methods, the most common of which use water, mud, or casing. Water stabilization is the least effective method, and works by reducing the stress level decreases on the sides of the hole. Further benefits come from the elimination of groundwater flow into the sides of the borehole. Water stabilization may work well in soft cohesive alluvial deposits, but it is not successful

13

Sampling and Sample Disturbance in a wide range of ground conditions. In partially saturated soils the loss of strength may encourage collapse, and in stiff fissured cohesive soils above the water-table the rate of swelling will be increased. Drilling mud may be made by mixing bentonite and water in a grout mixer, typically in proportions of about 1:20 by weight. Mud has several advantages over water. It has a higher density and therefore replaces a greater proportion of the stresses originally on the soil. It forms a ‘cake’ over any surface into which it attempts to seep; this cake is relatively impervious, thus reducing the rate and amount of swelling that can occur. The main disadvantage of mud is its high cost, and there are also problems with its disposal. For these reasons its use is normally restricted to rotary drilling. The most common method of stabilizing the sides of a borehole is the use of steel casing. Casing has the major advantages of being durable and providing a certain way of preventing collapse. Two types of casing coupling are in common use; the outside coupling and the flush coupling. The flush coupling is to be preferred because it minimizes the decrease in diameter of hole between adjacent casing strings and also because it suffers less from friction with the soil, thus allowing greater ease of extraction at the end of drilling a hole. Casing is usually advanced by driving with a heavy weight, such as a sinker bar or hammer. The use of casing can lead to certain types of soil disturbance, such as displacement, compaction, local overstressing and piping. Alternative effects of caving or collapse of the sides of the borehole however, can be equally severe and difficult to control without casing. Material which falls to the bottom of the hole shortly before an open drive sampler tube is lowered may be sampled and erroneously thought to be representative of soil conditions at that level. Immediately before any sampling is attempted, the depth of the base of the hole should be checked with a weighted tape to ensure that no debris has collected. If the depth of the hole is not equal to the last depth of the drilling tool, the borehole should be cleaned out and its depth checked once more, before a sampler is lowered. Small amounts of debris should be expected at the bottom of a borehole, but its depth should preferably never exceed 100 mm. Casing is normally fitted with a sharpened edge, or ‘shoe’ at its base. To minimize disturbance to surrounding soil this shoe should be kept sharp and should have an outside cutting edge. This will ensure that the soil displaced by the casing will be pushed into the borehole, from where it can be removed (Hvorslev 1940). Soil disturbance during sampling

Each type of sampling will impose a different degree and form of sampling disturbance, but in principle sampling processes can be divided into three broad groups. 1. Disturbed sampling. Here there is no attempt to retain the physical integrity of the soil. These types of sample are suitable for classification tests. 2. Tube sampling. The soil sample is obtained by pushing or hammering a tube into the ground. Soil is displaced and distorted, to a greater or lesser degree, as the tube enters the ground. There will be stress relief during boring, and during sampling when inside clearance is used. The design of the tube has an important effect on the disturbance of the soil. Tube sampling has, for the past 50 years, been the routine method of obtaining ‘undisturbed’ samples. 3. Block sampling. The sample is cut from the ground, either from the base or side of a trial pit, or as part of a rotary drilling process. Traditionally block samples have been obtained from pits. Carefully controlled rotary drilling, or the use of the Sherbrooke sampler, aims to achieve a similar result. Block samples undergo stress relief, and swelling, but should not be subjected to shear distortions. This section considers only block sampling and tube sampling. 14

Site Investigation

Block sampling Block sampling has traditionally involved the careful hand excavation of soil around the sample position, and the trimming of a regular-shaped block. This block is then sealed with layers of muslin, wax and clingfilm, before being encased in a rigid container, and cut from the ground. The process is illustrated in Fig. 6.5. A similar process can be carried out in shafts and large-diameter auger holes.

Fig. 6.5 Block sampling in a trial pit.

Trial pits are normally only dug to shallow depths, and shafts and large-diameter auger holes tend to be expensive. Therefore block samples have not traditionally been available for testing from deep deposits of clay. In the past decade, however, there has been an increasing use of rotary coring methods to obtain such samples. When carried out carefully, without displacing the soil, rotary coring

15

Sampling and Sample Disturbance is capable of producing very good quality samples. When the blocks are cut by hand then obviously the pit will be air-filled, but when carried out in a borehole it will typically be full of drilling mud. During the sampling process there is stress relief. At one stage or another the block of soil will normally experience zero total stress. This will lead to a large reduction in the pore pressures in the block. The soil forming the block will attempt to suck in water from its surroundings, during sampling, either from the soil to which it is attached, or from any fluid in the pit or borehole. This will result in a reduction in the effective stress in the block. In addition, where block sampling occurs in air, negative pore pressures may lead to cavitation in any silt or sand layers which are in the sample. Cavitation in silt and sand layers releases water to be imbibed by the surrounding clay, and the effect will be a reduction in the average effective stress of the block. Block sampling is an excellent method of ensuring that the soil remains unaffected by shear distortions during sampling, but samples obtained in this way may not (as a result of swelling) have effective stresses that are the same as those in the ground. Therefore the strength and compressibility of the soil may be changed. This should be allowed for either by using appropriate reconsolidation procedures, or by normalizing strength and stiffness, where appropriate, with effective stress. Tube sampling Tube sampling is used in almost all routine ground investigations. It is carried out by pushing a tube into the ground, without rotation, thus displacing soil. This displacement introduces shear distortions into the ground, and these can have two effects: 1. the effective stress of the soil is changed; and 2. bonding between soil particles (termed ‘structuring’) is broken. These effects are in addition to those induced by stress relief and swelling, described above for block samples, which occur in tube samples as a result of borehole disturbance and the design of the sampler. Baligh (1985), Chin (1986), Baligh et al. (1987), Siddique (1990), Hajj (1990) and Hopper (1992) have studied the penetration of samplers as a continuous flow problem. Early work by Baligh and his co-workers showed (Fig. 6.6) that the strains imposed on the centreline of a soil sample as it travels into a sample tube are initially compressive, and then extensive. The magnitude of the strains for the simple tube geometry that they simulated (the so-called ‘simple sampler’) depended on the thickness to widthratio (B/it) of the sampler. La Rochelle et al. (1987) believed that the idealization of tube geometry used in these early, pioneering works is not realistic, and stated that there is strong evidence that the detailed geometry of the cutting shoe of a tube sampler has a very large influence on the quality of sample obtained. Subsequent work by Siddique (1990) and Hopper (1992) has shown this to be true. Flat-ended samplers (Siddique 1990) and the simple sampler (Baligh et al. 1987) represent the extremes of poor design, and good cutting shoe design can very greatly reduce tube sampling disturbance, by reducing the magnitude of shear strains applied to the soil. Baligh et al. (1987) and Siddique (1990) applied the undrained strain paths deduced from Baligh’s strain path method to reconstituted normally consolidated unaged laboratory specimens, and observed the resulting stress paths. Hajj (1990) carried out tests on normally consolidated and overconsolidated reconstituted kaolin. Hopper (1992) carried out similar work on reconstituted overconsolidated unaged clay (OCR = 3.7), and in addition tested high-quality (Sherbrooke and Laval) samples of intact lightly overconsolidated estuarine clay. These tests have shown that for normally and lightly overconsolidated soils the stress paths during tube sampling are of the form shown in Fig. 6.7. Only very small strains are necessary to cause severe disturbance to unaged reconstituted normally consolidated clays. Lightly overconsolidated (OCR = 1.5) structured natural clays appear to be able to withstand axial strain

16

Site Investigation excursions of up to ±0.5% without significant loss of structure. Both normally and lightly overconsolidated soils suffer very large decreases in mean effective stress during tube sampling. More heavily overconsolidated clays appear to suffer little in the way either of destructuring or of effective stress change.

Fig. 6.6 Axial strain history at the centre-line of a simple sampler (after Baligh 1985).

On the other hand, tube sampling of heavily overconsolidated clays will often induce distortions of the type shown in Fig. 6.8. At the periphery of the sample the strains are similar to those imposed during simple shear testing; this cannot be modelled in the triaxial apparatus, where limited rupture zones occur at failure. This type of large-scale shear distortion results in a decrease in pore pressures, and an increase in the effective stress, in the periphery of the sample, which undergoes most shear distortion (Apted 1977; Hight 1986). If water is not available during the sampling process, either because the borehole is dry or because sampling takes places rapidly (and immediately after drilling to the required sampling depth), then pore pressure equilibration leads to a gradual increase in the mean effective stress in the centre of the sample, and a consequent increase in the undrained shear strength measured in laboratory triaxial tests.

17

Sampling and Sample Disturbance

Fig. 6.7 Stress paths induced by tube sampling on normally consolidated, and lightly overconsolidated clays

Fig. 6.8 Sketch of shear distortions induced in laminated heavily overconsolidated London clay by tube sampling.

Figure 6.9a shows estimates of the effects of tube sampling on the mean effective stress of London clay, shown by comparing the effective stresses in U100 tube samples (Fig. 6.11) with those in block samples (Fig. 6.5) from a similar depth in the London clay (about 22m below ground level) (Chandler

18

Site Investigation et al. 1992). The mean effective stress is apparently almost twice as large in the tube samples as in the block samples. This implies that the stiffnesses and strengths of tube samples will also be much larger, and this is shown by Fig. 6.9b. Here Hight (1986) shows an estimated profile of the in situ undrained shear strength, back-calculated from undrained triaxial tests and the mean effective stress in the ground, and indicates the magnitude of correction that should be applied to the uncorrected undrained strength obtained from tube samples in order to allow for the increase in shear strength due to tubesampling distortions. As might be expected from Chandler et al.’s findings, uncorrected strengths are about twice the estimated in situ values.

Fig. 6.9 Increases in effective stress in London Clay induced by tube sampling (Chandler, Harwood and Skinner 1992), and their effect on undrained shear strength (Hight 1986).

19

Sampling and Sample Disturbance Whilst the data in Fig. 6.9 indicate that strength and stiffness values obtained for stiff clays from tube samples should be reduced significantly, this cannot always be relied upon. Some clays contain significant laminations of silt or fine sand, or silt- covered fissure planes. Here sampling in the presence of water will probably be accompanied by swelling, since rapid penetration is possible. In other cases known to the authors, drilling fluid has not been completely removed from the top of tube samples, and mean effective stresses have been very significantly reduced. Also, swelling resulting from drilling disturbance may affect soil samples, especially when the base of the hole is not cleaned out immediately before sampling. As is evident from Fig. 6.8, significant shear stresses may be set up between sampling tube and the soil during driving. If the sampler is not properly designed these shear stresses can become sufficiently large that they prevent the entry of further soil into the sampler tube. This is termed ‘sample jamming’. Other pressures that may be applied to the soil during sampling include: 1. 2.

pressure on top of the sample, due to trapped borehole fluid, as the soil enters the tube; and tension at the base of the sample, as the tube is withdrawn from the base of the borehole.

Over the past half century practical experience has led to the development of an empirical design basis for samplers. This has been based upon the following parameters: • • • •

area ratio; cutting edge taper angle; lid ratio; and inside clearance.

In addition, it has long been known (for example, Hvorslev (1949)) that sample driving methods have a significant effect on the quality of the sample that is recovered. AREA RATIO Hvorslev (1949) defined one of the critical parameters affecting the disturbance of soil during sampling as the area ratio, defined by (see Fig. 6.10): D e − Di 2

area ratio =

De

2

2

(6.9)

where De = external diameter of the sampler cutting edge and Di = internal diameter of the sampler cutting edge. BS Code of Practice 2001:1957 specified that the maximum area ratio for the British Standard opendrive sampler should be 25%. The revised Code of Practice on Site Investigations (BS 5930) specifies a typical open-drive sampler as having an area ratio of ‘about 30%’. In view of the fact that Hvorslev (1949) noted that the incremental ratio of sample length recovered to length of drive was 1.25 (i.e. a greater length of sample was being obtained than the distance the tube was driven) for area ratios of 40—45%, this change seems retrogressive, since cutting edge taper angle is not specified. CUTTING EDGE TAPER ANGLE Increasing area (or kerf) ratio gives increased soil disturbance and remoulding, increased penetration resistance and the possibility of the entrance of excess soil from the area immediately beneath the cutting edge during the initial part of the sampler penetration. the permissible area ratio will depend on the soil type, its strength and sensitivity, and the purpose of sampling.

20

Site Investigation The use of very small area ratios leads to very fragile sampler tubes which may bend or buckle during driving the sampler into the soil. The practical need for a large ratio can be compensated for by the use of a small cutting edge taper angle, as proposed by the International Society for Soil Mechanics and Foundation Engineering’s Subcommittee on Problems and Practices of Soil Sampling (1965). For samplers of about 75 mm dia. they suggested the combinations of area ratio and cutting edge taper given in Table 6.10.

Fig. 6.10 Definition of area ratio and inside clearance.

It was also suggested that for clays the extreme edge of the cutting shoe could be given a 60° taper, until a 0.3 mm thickness was reached. In granular soils this thickness was suggested as the 10% grain size of the soil. When cutting edge taper angles are small, Scandinavian experience has shown that area ratios are largely irrelevant (Kallstenius 1958; Swedish Committee on Piston Sampling 1961). Table 6.10 Combinations of area ratios and cutting edge taper Area ratio Cutting edge taper (%) (deg.) 5 15 10 12 20 9 40 5 80 4

INSIDE CLEARANCE AND L/D RATIO With even moderate lengths of sample the adhesion or friction of the soil on the inside of the sampler tube may be sufficient to prevent further soil entering the tube. When wall friction is low it may produce slight compaction or compression of the soil, together with a down-dragging of soil layers at the edge of the samples. Severe distortion produces parabolic shapes in soil layers which are difficult to distinguish from plastic flow into the base of the borehole because of stress relief in soft soils. Severe effects of wall friction are transmitted to the soil lying beneath the bottom of the sampler.

21

Sampling and Sample Disturbance Ultimately, when the friction is large enough to prevent further entry of the soil into the sampler tube, bearing capacity failure of the soil beneath the bottom of the tube will take place. The soil will be severely remoulded and any material which enters the sampler will be useless even for visual examination. If samples are being taken continuously the top of the next sample will be worthless. One of the major factors controlling sample jamming is the length to diameter ratio of the sampler. The adhesion between a cohesive soil and the inside of a sampler barrel will be: A = πDLαc u

(6.10)

where D = inside diameter of the sampler, L = length of the tube, and α= reduction factor applied to the shear strength, cu, to give the adhesion between the soil and the tube. The bearing capacity of the soil beneath the tube is: q f = N c cu + p 0

(6.11)

where Nc is the bearing capacity factor (5—9), and p0 is the over-burden pressure and may approach zero if the borehole is large relative to the sampler. Equating the adhesive force to the bearing resistance of the soil, and taking p0 equal to zero, leads to:

L 1 Nc = D 4 α

(6.12)

In an extreme condition, taking α= 0.5 and Nc = 5, it appears that a maximum permissible length to diameter ratio of 2.5 should be considered. Three methods exist to reduce or eliminate wall friction between soil and the sampler; inside smoothness, inside clearance and sliding liners. The inside of all sampler tubes should be kept clean and smooth, and preferably polished. Oil may have to be used on old steel tubes, but this is not desirable. Inside clearance (see Fig. 6.10) is defined as the ratio:

inside clearance =

Ds − Di Di

(6.13)

where Ds = inside diameter of the sampler tube, and Di = inside diameter of the cutting shoe. Inside clearance gives the soil sample room for some swelling and lateral strain Jwing to horizontal stress reduction. Although neither of these types of behaviour is desirable, they are less undesirable than the consequences of adhesion between the soil and the inside of the sample tube. Inside clearance is usually less than 4%, because it should be large enough to allow partial swelling and lateral stress reduction but it should not allow excessive soil swelling or the loss of the sample when withdrawing the sample tube. Hvorslev (1949) suggests 0.75—1.5% inside clearance for long samplers, and up to 1.5% for very short samplers: he suggests an inside clearance of between 0.75 and 1.5% under average conditions. Where inside clearance of this magnitude is provided, Hvorslev recommends that for ‘properly designed and operated’ drive samplers of 50—75mm inside diameter the maximum length to diameter ratios can be increased as follows: loose to dense cohesionless soils L/D>|5—10, and very soft to stiff cohesive soils L/D>|10—20.

22

Site Investigation The ISSMEFE Report of the Subcommittee on Problems and Practices in Soil Sampling (1965) suggested that if the inside surfaces of the sampler tube are smooth and clean and the coefficient of friction is low, an inside clearance of 0.5—1.0% is suitable for sampling to depths of 20m in ‘nonswelling’ soils. With this inside clearance the permissible length to diameter ratio should depend on the soil type as shown in Table 6.11. Table 6.11 Dependence of permissible length to diameter ratio on soil type Type of soil Greatest length to diameter ratio Clay (sensitivity> 30) 20 Clay (sensitivity 5—30) 12 Clay (sensitivity < 5) 10 Loose frictional soil 12 Medium loose (?) frictional soil 6

The Subcommittee commented that large inside clearances (>1—3%) cause deformations of the samples, opening of fissures, and swelling of soils containing gases: ‘A need for excessive inside clearances may indicate bad sampler design or sampling technique’. Inside clearance has always been regarded as a necessary evil, and recently some samplers have been designed which deliberately do not make use of it. These samplers (for example, the Laval sampler — see Chapter 7) are intended for use in normally and lightly overconsolidated and sensitive clays, where disturbance at the sample periphery will produce a very low-strength clay, that is to some extent selflubricating. They have low length to diameter ratios. The use of zero inside clearance to sample heavily overconsolidated clays cannot be recommended unless the length to diameter ratio of the sample can be less than 2. The use of sliding liners inside sampler tubes would appear to be preferable to the use of inside clearance. Samplers described by Kjellman et al. (1950) and Begemann (1961) use foil and stockinette respectively, and can give near continuous samples of great length. The disadvantages of these types of sampler lie in their great cost. SAMPLE DRIVING METHODS Sample driving methods can have a severely damaging effect on soil. The effects of trying to drive a thick walled open-drive sampler into hard soil by repeated blows of a hammer are obvious; the soil is usually heavily fractured and if any material is recovered it often has the appearance of an angular gravel. The method of driving a sampler is often crucial, not only to the disturbance of the soil, but in consequence to the ability of a sampler to recover it. Hvorslev (1949) rates drive methods as shown in Table 6.12. Table 6.12 Driving methods (Hvorslev 1949)

Method

Motion

Hammering: repeated blows of a drop hammer Jacking: levers or short commercial jacks Pushing: steady force — no interruptions

Intermittent fast motion Intermittent slow motion Continuous uniform motion Continuous fast motion Continuous very fast motion

Single blow: blow of a heavy drop hammer Shooting: force supplied by explosives

Sample quality Worst

Best

23

Sampling and Sample Disturbance Hammering is a method commonly used to advance open-drive samplers into the ground, particularly in conjunction with light percussion drilling. The hammering action may take place down the hole, or at the top of the hole. In the former method (Fig. 6.11) the sampler tube is separated from a weight (the sinker bar) by a jarring link. The sampler tube is advanced into the soil by repeatedly lifting the sinker bar and allowing it to fall on the drive head. The use of a relatively thin and sometimes worn jarring link at the base of the borehole allows the sampler tube to rock from side to side; this can lead to breaks in the sample. Similarly severe effects can be produced if the sinker bar is lifted too high during driving, when the sampler tube will be pulled upwards and tension applied to what will be the middle of the sample.

Fig. 6.11 U100 sampler assembly and details of cutting shoes.

If samples are to be hammered into the soil, then it is essential that the sampler should be rigidly 24

Site Investigation connected to rods extending to ground level. If the borehole is deep and large compared with the rod size, spacers may be required to reduce rod buckling as the hammer energy travels to the base of the hole. Hammering is cheap, but gives poor quality samples. At the other end of the scale, a single blow or the use of explosives will give a relatively high energy input which is difficult to control. One of the obvious dangers is that the sampler will be driven too far, leading to compaction of the material within it. The best practical method of sample driving is therefore pushing. Most modern auger rigs can readily supply a steady downwards force, with no interruptions, but a light percussion rig will need some adaptation. A typical arrangement for pushing a piston sampler into soft ground is shown in Fig. 6.12.

Fig. 6.12 Continuous push driving by means of winch and block and tackle.

When driving sample tubes into the ground, by whatever method, it is important to remember that water (or air) above the top of the sample or piston, contained inside the tube, must be able to escape without significant increases in pressure occurring. It is normal to provide vents in the top of the sampler, but their size must be limited for reasons of geometry and sampler strength. Therefore it is necessary to limit the speed of sampler penetration. For most samplers a speed of 25mm/s will be satisfactory.

Disturbance after sampling

Changes to the soil after sampling can be at least as severe as those occurring during boring and sampling. Five major types of change can be recognized: 1. moisture loss; 2. migration of moisture within samples;

25

Sampling and Sample Disturbance 3. the effects of inadvertent freezing; 4. the effects of vibration and shock; and 5. the effects of chemical reactions. Moisture loss Representative samples do not need to have their moisture content preserved, but it is often helpful to the engineer if considerable moisture loss is not allowed to occur. In order to restrict moisture loss and prevent loss of fine soil particles it is normal, therefore, to place the soil in heavy gauge polythene bags, boxes or glass jars. Block samples and tube samples must not be allowed to lose moisture. Hvorslev (1949) reports the results of long-term experiments with different sealing methods which are shown plotted in Fig. 6.13. These results indicate that the best sealant for tube samples was battery sealing compound, with a water loss of only 0.1 g after sample storage in a tube with 3/4in. (19mm) sealing compound plugs. The stickiness of this asphaltic material, however, makes subsequent removal from soil and cleaning of tubes very difficult, and it is now rarely used.

Fig. 6.13 Moisture loss with various sample sealing methods (data from Hvorslev 1949).

Block samples should be sealed by initially applying a brush coating of 2mm of paraffin wax, followed by wrapping in cheesecloth or clingfilm and subsequent dipping to increase the covering to at least 4mm. Large block samples which cannot be readily dipped can be placed in an oversize box and encased in paraffin wax poured around the soil to fill the void. Alternatively, after an initial brush coating of wax, the sample can be wrapped in aluminium or clingfilm, waxed again, and placed in a wooden box and encased in polyurethane foam. Paraffin wax shrinks upon cooling, and small cracks will lead to rapid moisture loss. All paraffin wax should be applied as close to melting point as possible to reduce shrinkage. At this temperature it

26

Site Investigation should be possible to dip a finger in the wax without being burnt. (Note that the melting point of paraffin wax is about 50°C, and therefore in very hot climates it will not be a suitable sealant.) Tube samples are generally sealed with paraffin wax, but Hvorslev’s experiments showed that even when properly applied this material will tend to undergo plastic deformations after about six months. Once defects appear in the seal, moisture loss is rapid. The best sealing method for tubes appears to be the use of tightly fitting plastic caps, but when these are applied to the sample the necessary escape of air may leave an unsealed channel which will allow moisture loss. These channels can often be closed by rotating the cap once in place, or by using plastic self-adhesive tape. Alternatively, equally effective sealing methods are vented push-on or screw-on caps, combined with paraffin wax seals. With 0.4 mm paraffin wax and vented caps Hvorslev (1949) observed a 6.8g moisture loss in 1250 days. In this type of sealing it is important to fill the gap between the sample and the caps if the sample is short, since this will reduce the loss of moisture into the airspace and will also stop the sample sliding in the tube during handling. Where caps are not available, the wax seals should be reinforced with metal discs, made of a material such as aluminium foil. This will reduce shrinkage and eliminate the formation of pinholes in the central section of the end of the sample. If wax is used to seal samples of high permeability, or where voids or discontinuities are open in the soil, the penetration of wax poured directly on to the end of such a material may reduce its worth. These sorts of problem can be overcome by brush application of the first layer of wax, followed either by application and wax soaking of a layer of cheesecloth or by application of a layer of aluminium foil and further paraffin wax layers. Even after the samples reach the laboratory. care must be taken. Storage conditions are important. As seen above, wax seals will often become ineffective after only a few months. Where sealing has been carried out to a high standard, the speed at which seals deteriorate is increased by a high storage temperature. Samples should therefore be stored in a cool room, with temperatures preferably not exceeding 30°C. Low temperatures and high humidity will help to reduce moisture losses once imperfections appear in the wax. Migration of moisture within samples Once samples are adequately sealed, migration of water within the sample may still lead to significant changes of properties such as undrained strength and compressibility. Two types of effect have been noted; in the first, water migrates from one type of soil to another (Kimball 1936; Rowe 1972); while in the second, differential residual pore pressures in the samples equalize with time (Casagrande 1936; Schjetne 1971). Consider a laminated soil, containing alternate layers of silt grading into fine sand and clay. In situ the clay might have a firm or stiff consistency, but once stress relief occurs the water in the granular layers will migrate to the clay and relieve the negative excess pore water pressures. Upon examination, the soil might appear to consist of very soft clay layers interbedded with relatively dry silty sands. As an example of the second type of moisture migration, consider the effects of sampling on a very stiff clay of high plasticity, such as the London clay. After sampling the bulk, the soil (its inner portion) will be expected, as a result of stress relief, to have strongly negative pore pressures, whilst retaining effective stresses similar to those that it had in the ground. As a result of the type of shear distortions induced by tube sampling (Fig. 6.8) the outer part of the sample will have lower pore pressures, but if rotary coring has been used, or the sample tube was left in a water- filled borehole, then there may be have also been some swelling, and a consequent increase in the pore pressures around the outside of the sample. With time, as pore pressures equalize in the sample, there will be a change in the average effective stress in the sample, and therefore a change in the strength and compressibility that will be measured in the laboratory. Hight’s results (Fig. 6.9) and those of Apted

27

Sampling and Sample Disturbance (1977) suggest that generally tube sampling produces an increase in measured unconsolidated undrained strength (cu). Rotary coring leads to a decrease in measured undrained strength. The change in strength and stiffness that is measured will, to some extent, be time dependent, because water must flow between the outside and the centre of the sample in order that pore pressures may equalize. Shear strengths measured immediately after sampling will be different from those measured, for example, after the samples have been transported to a laboratory and stored for some time. In soft clays, Casagrande (1936) noted that the outer layer of a soil sample would have higher pore pressures than the centre immediately after sampling, as a result of the higher tube sampling strains that are experienced. This has been confirmed by a number of researchers (Schjetne 1971; Bjerrum 1973; Siddique 1990; Hopper 1992). Schjetne (1971) has measured the effects of sampling soft clays of different sensitivities with a Norwegian Geotechnical Institute piston sampler on pore pressures within the sample during sampling and after extrusion. His observations confirm Casagrande’s mechanism. Bjerrum (1973) has shown that owing to remoulding and moisture migration, the outer 5mm of extruded Drammen clay specimens typically have a moisture content about 3—4% lower than at the centre. To avoid this Casagrande recommended that the outer disturbed layer of the soil samples should be shaved off as soon as the samples are removed from the borehole. Freezing Probably the most serious effects of poor storage will occur if clay or silt samples are allowed to freeze (Kallstenius 1958). Ice lenses form initially in fissures, and the soil is gradually broken up by a wedging action as water is attracted from the rest of the sample to these lenses. Frozen samples are highly disturbed samples, and therefore a sample store should never be allowed to drop to temperatures below 4°C.

Vibration, shock, and mechanical disturbance Vibrations caused during the transportation of some soils to the laboratory may cause a loss of strength and remoulding (Kallstenius 1963) particularly on very soft silty or sandy clays stored in horizontally positioned tubes. Compaction effects may cause soil distortion, the liberation of pore water and the movement or break-up of the wax seals at the ends of the specimen. Preventing these effects is often rather difficult, but they can be reduced by supporting the samples vertically on a compressible base such as a foam mattress. Because they have no supporting tube, block samples of very soft, soft and sensitive clays are often at considerable risk from sudden shocks during handling and transporting to the laboratory. It is therefore advisable to place each sample in a separate rigid container, surrounding it with packing material to prevent it from moving. Loose granular samples are likely to undergo density changes even when very carefully handled. In the last stage of the life of a sample it will be extruded. The rules for good extrusion should be based on the same factors as control sampling. The soil should be pushed out at a steady speed. To avoid disturbance and distortion of the soil layers a plunger of almost the same diameter as the inside sampler diameter should be applied to the bottom of the sample, so that the same relative movement between soil and tube is continued. This means that the top of the sample must be marked in the field, and the first soil to emerge will be disturbed and should be discarded. Various methods are available to provide force to an extruder plunger; these include direct fluid pressure, hydraulically operated pistons and mechanical devices. Air or water will often penetrate past a plunger, and can cause considerable disturbance to the sample. The most reliable systems use either a continuously screw-threaded shaft or an hydraulic piston to advance the plunger. Of those two

28

Site Investigation methods, the hydraulic system is the most convenient, provided that it can provide a stroke of sufficient length. Reactions between soil and tube during storage Since it may be necessary to store samples for some time before laboratory testing can be carried out, there may be a considerable opportunity for chemical reaction between the soil and the sampler tube. Acid and alkali soils will attack sampler tubes, as will soil specimens with saline pore water. Further problems may occur if the tube and end cap are made of different types of metal. Changes in the pore water chemistry can have serious effects on soil behaviour, for example decreasing sensitivity. Electrolytic action may cause a change in soil plasticity, compressibility and shear strength. Disturbance in the soil-testing laboratory

Even when the utmost care is taken to avoid the serious effects that have been described above, it is still possible for soil testing to be carried out on disturbed materials, as a result of further disturbance induced once the sample enters the laboratory. The principal causes of disturbance are: 1. poor extrusion practice, either due to high extrusion pressures being applied to unsaturated soil, or due to lack of proper support of low-strength clays during extrusion; 2. use of poorly designed tubes to take small-diameter specimens from larger diameter samples; and 3. damage to soil ‘structure’ as a result of poor saturation or reconsolidation procedures. Effects of sample disturbance

The most obvious effect of sample disturbance can be seen when attempting to tube sample very soft, sensitive clays with a poorly designed sampler. The soil around the edge of the sample undergoes a very large decrease in strength, such that when the tube is withdrawn from the soil there is no recovery. But, as has been noted above, sample disturbance occurs in all sampling processes and, if sampling is carried out well, the effects of disturbance will hopefully be more subtle. Whatever its magnitude, sampling disturbance normally affects both undrained strength and compressibility. In addition, chemical effects may cause changes in the plasticity and sensitivity of the soil sample. Failure to recover Failure to recover is the most serious result of sample disturbance and can be caused by a number of factors such as: 1. Remoulding adjacent to the sampler walls. Adhesion or friction is required to support the soil when the sampler is being lifted out of the ground. Many soils exhibit sensitivity (i.e. a loss of shear strength during remoulding), and the remoulding of soil adjacent to the sampler barrel therefore reduces the chances of recovery. In soft or very soft soils a low area ratio or cutting edge taper angle is essential. 2. Pressure over the top of the soil sample can be created if no vent is built in to allow air or water to escape as the soil enters the lower end of the sample tube, or if the vent is too small and the velocity of soil entry is large. When pulling the soil samples from the soil, pressure over the sample will help to push it out of the tube. Most samplers are provided with one or more vents in the head. It is essential that they be kept clean. 3. Suction beneath the sample will occur as the sample tube is pulled from the soil, since a void must be created at the level of the base of the sample. This effect can be reduced or eliminated by either fixing lugs to the sides of the tube which will give an outside clearance if the sample

29

Sampling and Sample Disturbance tube is rotated (Harper 1931) or by providing pipes down or in the sides of the sampler tube to allow injection of air or water at the base of the sample (Mohr 1943). Alternatively, suction may be opposed by using a piston sampler; if the soil tries to slide out of the base of the tube then a suction force will also be set up at the top of the sample. The use of suction at the top of a sample is apparently incorporated into many open-drive sampler tubes by the use of a ‘ball valve’ in the head (see Chapter 7). Even when perfectly clean however, the ball will not normally seat perfectly to provide an efficient seal and prevent re-entry of air into the top of the tube if the soil should start to fall out. 4. The tensile strength of the soil at the base of the sampler must be overcome. If the sampler is simply pulled vertically then the combination of disturbance, vacuum and tensile strength will often be sufficient to cause loss of recovery. To overcome the tensile strength the sampler may be rotated two or three times before being gently pulled upwards. Rotation of the sampler will induce torsional soil failure at the base of the cutting shoe. More sophisticated and less practical methods have been used involving snare wires (Buchanon 1936, 1938; Hvorslev 1940) or pushed curved springs which cut the sample free. These are unnecessary for routine work. When soil samples are lost a number of simple techniques can be tried to improve recovery. These include the following: i. A rest period after driving the sampler and before extracting it will allow the soil to swell inside the sample tube, improving the adhesion of fatty overconsolidated clays to the side of the tube. ii. Slight over-driving, which also increases soil disturbance, will help the retention of both cohesive and non-cohesive soils since it will splay them against the side of the tube and improve friction or adhesion. iii. Core retainers (core catchers, catcher boxes) can be incorporated in the cutting shoes of open-drive samplers to improve recovery. The most common designs are the ‘basket’, a series of curved springs mounted in or immediately above the cutting shoe, and the use of hinged flaps mounted in the upper part of the cutting shoe. Core retainers often cause severe disturbance around the edge of the sample, and the sampler area ratio will need to be large to accommodate them. Strength Although it has been noted above that tube sampling disturbance has the greatest effect, in terms of reductions in mean effective stress, on reconstituted clays its effect on the undrained shear strength of such material is, perhaps surprisingly, small. Laboratory experiments by a number of workers have shown that the stress paths during undrained shearing converge on the critical state and, because the soil is initially reconstituted, the state boundary surface is not disrupted by tube sampling. Typically, it has been found that the undrained strength is reduced by less than 10%, even when the material is not reconsolidated back to its initial stress state (for example, Siddique (1990)). Tube sampling does, however, have a significant effect on real soils, most of which are either bonded (‘structured’), and/or more heavily overconsolidated. Shearing of bonded soils during tube sampling can have the effect of progressively destructuring them. Clayton et al. (1992) show comparisons of the stress paths taken by soil specimens tube sampled in different ways. Figure 6.14 shows how tube sampling a lightly overconsolidated natural, structured clay with a standard piston sampler leads subsequently to much higher pore pressure generation during undrained shear, with the consequence that undrained strength is reduced. Clayton et al. (1992) found that provided tube sampling strain excursions were limited to ± 2% and that appropriate stress paths were used to reconsolidate the material back to its in situ stress state, the undrained strength of the Bothkennar clay would be within ± 10% of its undisturbed value. It is to be expected, however, that much greater effects will occur when sensitive clays are sampled.

30

Site Investigation

Fig. 6.14 Effects of tube sampling disturbance of lightly overconsolidated natural (‘structured’) clay on: (a) stress path and strength during undrained triaxial compression (b) one-dimensional compressibility during oedometer testing.

Heavily overconsolidated clays often display almost vertical stress paths under undrained shear. An increase in the mean effective stress level as a result of tube sampling will result in approximately proportional increase in intact strength. Unfortunately, however, this is not the only effect at work. Hammering of tubes into stiff clays can cause fracturing, and loosening along fissures, and this may lead to a marked reduction in measured undrained strength. In a simple study of the influence of different methods of sampling, Seko and Tobe (1977) measured the unconfined compressive strength as a function of depth obtained from samples taken using different sampling devices. The very wide variation in the strength of stiff Tokyo clay can be seen in Fig. 6.15, which shows that thin-walled open-drive hammered tube sampling gave much lower strengths than double-tube rotary coring methods with mud flush — the opposite of what might be expected from simple considerations of effective stress change alone. Single-tube rotary coring with a tungsten bit produced the lowest strengths.

31

Sampling and Sample Disturbance

Fig. 6.15 Variation in unconfined compressive strength of Tokyo clay caused by different sampling methods (Seko and Tobe 1977).

Compressibility and stiffness The effects of sampling on compressibility (as measured in the oedometer, for example) are difficult to assess because of bedding effects, particularly in heavily overconsolidated clays. The use of local axial strain measurement on triaxial specimens during the past decade (see Chapter 8) has produced new and more reliable stiffness data than can normally be expected from routine one-dimensional consolidation tests, It is now known that the measured small-strain stiffnesses of clays, most relevant to many geotechnical engineering problems, is for a given clay approximately linearly proportional to the mean effective stress at the time of measurement. ,This means that changes in effective stress as a result of disturbance are directly translated into proportional changes in measured soil stiffness. Because of the growing appreciation of the influence of bedding and effective stress changes on measured stiffness, it has become common practice in the UK to adopt laboratory methods which will avoid these problems. In heavily overconsolidated clays, small-strain stiffness is often normalized with respect to the mean effective stress at the start of shear (p’o=(σ’1+σ’2+σ’3)/3). Alternatively, the stiffness of bonded soils is perhaps more appropriately normalized with respect to undrained shear strength, although it may be difficult to determine the true in situ value of this. In situ stiffness can then be recovered if p’o(in situ) or cu(in situ) be estimated. In lightly overconsolidated natural clay Clayton et al. (1992) have shown, however, that even the careful reestablishment of in situ effective stress levels before shearing cannot fully recover the undisturbed stiffness behaviour of the soil. A 60% reduction in Eu/ p’o (measured locally, and after re-establishment of in situ stresses) was obtained for the Bothkennar clay following tube sampling strain excursions of ±2%, for example. The results of a literature survey by Hopper (1992) are shown in Fig. 6.16. Here the very severe effects of tube sampling (including the effects of borehole disturbance, and obtained by comparing test results from tube samples with those on block samples in the same soil type) can be seen.

32

Site Investigation

Fig. 6.16 Influence of tube sampling disturbance on undrained strength and stiffness (from a survey by Hopper 1992).

Siddique (1990) carried out an analytical study of typical sampler cutting shoe geometries, and found that: 1. increased area ratio, as a result of increasing the thickness of the sampler tube (and therefore decreasing the B/t ratio (Fig. 6.6)) causes a significant increase in the peak compressive strain occurring ahead of the sampler, but has only a limited effect on the peak extensive strain; 2. increasing inside clearance as a result of increasing the inside diameter of the sampler tube causes a significant effect on the peak extensive strain, and a slight decrease in the peak compressive strain; and 3. outside cutting edge taper angle has a marked effect on the peak axial compressive strains experienced by a sample. In order to restrict the peak axial compressive strains (both in extension and in compression) to less than 1%, he recommends the following design for tube sampler cutting shoes: • • • •

area ratio >|10%; inside clearance ratio>|0.5%; inside cutting edge taper angle 1 to 1.5°; outside cutting edge taper angle >| 5°.

In addition, Hvorslev’s work indicates that tube samples should be pushed smoothly into the soil, in a single smooth action. Even given a maximum strain of 1%, normally consolidated reconstituted clays show considerable signs of disturbance. Compared with an ‘undisturbed’ specimen of reconstituted London clay, Siddique (1990) found the following reductions in effective stress, strength and stiffness: p’0 E50 (Eu)0.0l %/ p’0 Cu

26% 65% 78% 6%

Strain path tests on very high quality (Laval and Sherbrooke) undisturbed samples of natural clay by 33

Sampling and Sample Disturbance Hopper (1992) has confirmed that for normally and lightly over- consolidated clays, stiffness is greatly affected by tube sampling, but that undrained strength reductions are less significant and can, in any case, be recovered by good reconsolidation procedures. CLASSIFICATION OF SOIL SAMPLES

Hvorslev ‘s classification

Despite the more recent, and more sophisticated classifications which have been produced subsequently (see below), it is Hvorslev’s (1949) classification of soil samples which remains widely used in British ground investigation. It is simple, and in view of the fact that we must now recognize that all soil will undergo some disturbance before reaching the laboratory test apparatus, there is arguably no need to further subdivide his categories. Hvorslev considers only three classes of sample. 1. Non-representative samples are samples containing mixes of soil or rock from different layers, or soils where certain fractions have been removed or exchanged by washing or sedimentation. This type of sample is now not normally considered as useful in site investigation, particularly since considerable skill may be required even to obtain a preliminary classification of the sub-soil. This type of sample is typically produced by the following. i. Washboring — where progress is made by jetting, and tests are made on open drive samples, fine granular soils may be washed away, and coarse granular particles may collect at the base of the hole, giving false particle size distributions in samples. ii. Bailing — the use of a ‘shell’, ‘bailer’ or ‘sand pump’ while percussion drilling forces the soil at the base of a borehole into suspension in the water. The coarse fraction of the soil will tend to sediment quickly, while silt- or clay- size material will remain in suspension in the water, and will often either be left in the borehole or tipped away before samples are taken. iii. Rotary open-holing — which uses a similar technique to washboring to advance the hole. Gravel-size particles will not be lifted up the hole, except by unacceptably high up-hole flush velocities which will lead to excessive borehole erosion. 2. Representative samples are samples of soil from a particular stratum which have not been contaminated by minerals or particles from other levels in the borehole, and have not been chemically altered, but may have been remoulded and have had their moisture contents changed. These samples may be obtained from samplers which are unsuitable for the soil conditions, or where samples are taken from the cutting shoe of samplers before they are sealed. In addition representative disturbed samples may be obtained from material obtained from relatively uniform soils by claycutter, or where clay materials are removed from the sampler shortly after sampling and placed in containers which allow them to alter their moisture content with time. Hvorslev’s classification differs from Rowe’s in that Rowe terms the ideal sample as representative (i.e. moisture content, material content, fabric and structure and stress state all remain unaltered). Hvorslev’s ‘representative samples’ correspond to the British ‘disturbed samples’ which are sometimes specified as ‘to be truly representative of the composition of the in situ soil’. 3. Undisturbed samples are samples in which the soil is subjected to little enough disturbance to allow laboratory experiments to determine the approximate physical characteristics of the soil, such as strength, compressibility and permeability. Hvorslev’s ‘undisturbed samples’ correspond to Rowe’s quality class 1 and 2 because although quality class 3 utilizes driven or pushed thin- or thick-walled samplers, these may not be suitable for the soil conditions and may lead to sampling disturbance.

34

Site Investigation Rowe’s classification

The problem facing the engineer is to obtain adequate samples for the purposes envisaged. Rowe (1972) has defined five qualities of soil sample, based on the German work of Idel et al. (1969) (Table 6.13). This classification places heavy emphasis on the use of water balance. This means that where artesian conditions are encountered soil samples intended to be quality 1 or 2 must be taken from a rig mounted on a platform with casing extending above ground level, or by using drilling mud. Table 6.13 Sample quality classes after Idel et al. (1969) as modified by Rowe (1972) Quality Required soil properties Purpose Typical sampling class procedure Laboratory data on Piston thin-walled 1 Remoulded properties in situ soils sampler with water Fabric balance Water content Density and porosity Compressibility Effective strength parameters Total strength parameters Permeability* Coefficient of consolidation* Laboratory data on Pressed or driven thin2 Remoulded properties or thick-walled in situ insensitive Fabric sampler with water soils Water content balance Density and porosity Compressibility* Effective strength parameters* Total strength parameters* Fabric examination Pressed or driven thin3 Remoulded properties and laboratory data or thick-walled Fabric A* 100% recovery Continuous on remoulded soils samplers. Water balance in B* 90% recovery Consecutive highly permeable soils 4 Remoulded properties Laboratory data on Bulk and jar samples remoulded soils. Sequence of strata Washings 5 None Approximate sequence of strata only *Items changed from German classification.

BRITISH PRACTICE, AND THE BS 5930 CLASSIFICATION

British site investigation practice at present commonly divides samples into the following categories. 1. Disturbed samples: small disturbed samples (‘jars’); and i. ii. large disturbed samples (‘bulk bags’). 2. Undisturbed samples: i. block samples; ii. open-drive samples; piston-drive samples; and iii. iv. rotary core samples (such as from the corebarrel).

35

Sampling and Sample Disturbance

All of these samples are intended to be representative of the composition of the in-situ soil; nonrepresentative samples are, of course, never intentionally taken. It is important to recognize the distinction between sampling in cohesive and noncohesive soils. In cohesive or cemented soils it is usually possible to obtain what Hvorslev termed ‘practical undisturbed’ samples, and there is a wide variety of sampling equipment and laboratory equipment to obtain and test such samples. In non-cohesive soils Rowe (1972) has stated it is doubtful whether Quality 1 samples have even been obtained’. The problems of obtaining and testing non- cohesive samples can be briefly summarized as follows. 1. Volume changes during driving or subsequent handling of sampler tubes, due to vibrations. 2. The soil may collapse if unsupported. Special precautions must be taken to get the soil into the test apparatus without releasing compressive stresses. 3. High friction, developed as the sample enters the tube, may remould or alter the stress levels on the soil. 4. Inevitably some modification of the stress levels on the sample will take place. The strength and compressibility of non-cohesive materials are highly stress-dependent. In special circumstances, samples have been obtained using freezing (Fahlquist 1941) and chemical injection (Van Bruggen 1936; Karol 1970). Both these techniques alter the soil, the first by volume change and the second by contamination. Rowe claims that Quality 2 samples may be obtained using mud or water-filled boreholes and thin-walled pushed piston samplers have been successfully used in medium dense sands, but in most cases it will be sufficient only to consider obtaining Quality 3 samples to allow fabric examination for the planning of in situ tests. Quality 3 samples may often be obtained using relatively common sampling techniques, such as thin-wall piston sampling. One device specifically designed for sand sampling has been described by Bishop (1948).

36

Chapter 7

Undisturbed sampling techniques INTRODUCTION

Of the very large number of sampling techniques devised worldwide since the turn of the century, few are now in current use, and even fewer are in current use in the UK. Here, the most widely used tools are the 100 mm dia. thick-walled open-drive sampler, the ‘Standard Penetration Test’ 35 mm thickwalled open-drive split barrel sampler, 54mm or 102mm thin-walled fixed-piston samplers, and double-tube swivel type core barrels. The types of sampler adopted in each part of the world depend on the state of development of the area, its sampling tradition, economics, and its principal soil types. In the heavily developed South East and Midlands of England, soil types are typically stiff or very stiff clays and weak rocks. In the valleys, alluvium often consists of coarse gravels. Sampling is therefore based on the use of rugged tools in a large diameter borehole. When carrying out site investigation abroad, the available drilling equipment is often very different from that used at home, and the familiar sampling tools may be either unobtainable or inappropriate. When drilling at home the solution of new problems may require a reappraisal of the value of commonly used techniques. These factors require an engineer to be aware of as many types of sampler as possible and this chapter therefore sets out to review the main types of equipment now available. In Chapter 6 (Sampling and sample disturbance) the way in which a number of common types of sampler are constructed, and the manner in which they work was described. Samples are obtained in a number of ways: 1. by using a number of techniques in shallow pits, shafts and exposures; and 2. in boreholes, using either drive or rotary techniques. Drive samplers are pushed into the soil without rotation, displacing the soil as they penetrate. They generally have a sharp cutting edge at their base. In contrast, rotary samplers (often termed ‘corebarrels’) have a relatively thick and blunt cutting surface, which has hard inclusions of tungsten or diamond set into it. The sampler is rotated and pushed (relatively) gently downwards, cutting and grinding the soil away beneath it. A general classification of samplers is shown in Fig. 7.1.

Fig. 7.1 General classification of borehole sampling devices.

Undisturbed Sampling Techniques It is generally believed that undisturbed sampling is not possible in granular soils. Nonetheless, special techniques for sand sampling have been developed over the years, and these are described in a later section of this chapter. Finally, we consider the selection of appropriate samplers for different purposes, and to suit different ground conditions. SAMPLES FROM PITS AND EXPOSURES

Trial pits, trenches and shallow excavations are often used in site investigations, particularly during investigations for low- and medium-rise construction, because they provide an economical means of acquiring a very detailed record of the complex soil conditions which often exist near to the ground surface. It is worth remembering, however, that trial pits and other exposures can also be used for in situ testing and to obtain high-quality samples. The types of samples taken will vary according to the needs of the investigation. Disturbed samples of granular soil are likely to be more representative than those that can be taken from boreholes. Disturbed samples are often taken for moisture content or plasticity determination in the laboratory, and in association with determinations of in situ density. In situ density testing is described in Chapter 9. Undisturbed samples can be obtained either by drive sampling (see below) or block sampling (as described in Chapter 6). In either case it is important to recognize the disturbance created by excavating the trial excavation, and ensure that disturbed material is carefully removed before or after sampling. To this end, the faces and bottom of the pit should be hand trimmed in the areas to be sampled (or described), particularly when the pit has been machine excavated. In exposures, an attempt should be made to remove the weathered surface of the soil. In the UK 38mm open-drive tubes are often hammered into the sides and base of trial pits. These tubes normally have no check valve, a high area ratio, and no inside clearance. It is often necessary to dig the sample tube out of the soil in order to avoid losses. When U100 tubes (see under ‘Drive samplers’) are used, they require considerable force and are commonly pushed into the soil with a ‘back-actor’ bucket. Care must be taken to prevent rocking of sample tubes during driving since this causes serious disturbance to the soil at the shoe level. The use of a frame to align the sampler during driving is advisable. Better quality samples of firm to stiff clay soils can be obtained by trimming the soil in advance of a large diameter (100—200 mm) sampler. This eliminates the disturbance caused by soil displacement ahead of the cutting shoe, but may allow slight lateral expansion of the soil. When the material to be sampled is either hard, stoney or coarse and granular, it is essential that advanced trimming of large diameter samples is used. When the soil is sufficiently stiff or cemented to stand up under its own weight, a block sample may be taken. The normal technique is to cut a column of soil about 300mm cube, so that it will fit inside a box with a clearance of 10—20 mm on all sides. A box with a detachable lid and bottom is used for storage. With the lid and bottom removed, the sides of the box are slid over the prepared soil block, which is as yet attached to the bottom of the pit. After filling the space between the sides of the box and block with paraffin wax, and similarly sealing the top of the block, the lid is placed on the box. The block is then cut from the soil using a spade, and the base of the sample trimmed and sealed. Block samples allow complete stress relief, and may therefore lead to expansion of the soil, but in very stiff clays this technique is widely regarded as providing the best available samples (see Fig. 6.5).

2

Site Investigation The Sherbrooke sampler Block samples can only be taken from depth in heavily overconsolidated soils, such as the London clay. In normally and lightly overconsolidated clays, excavation of a pit or shaft to more than a few metres depth is often impossible because base heave will occur. Lefebvre and Poulin (1979) calculate that, for example, in a clay with an undrained shear strength the depth of a trench or pit will be limited to about 4 m, if a factor of safety of two is to be maintained. To overcome this problem, Levebvre and Poulin (1979) designed the apparatus shown in Fig. 7.2, which is essentially a down-borehole block sampler. The equipment needs a borehole of about 400 mm diameter, which is best cleaned using a flat- bottomed auger, in order to reduce disturbance and minimize the amount of disturbed material left in the base of the hole before sampling. The hole is kept full of bentonite mud. The sampler is lowered to the base of the hole, and rotated, either by hand or using a small electric motor, at about 5r.p.m. A cylinder of soil, about 250 mm in diameter, is carved out by three circumferential blades, spaced at 120°. They make a slot about 50mm wide, and are fed by bentonite or water to help clear the cuttings. The time taken to obtain a sample obviously depends upon ground conditions, but may be about 30—40min. After carving out a cylinder about 350 mm high, the operator pulls a pin, and the blades (which are spring-mounted) gradually rotate under the base of the sample, as rotation is continued. Closure of the blades separates the sample from the underlying soil, and the sample is then lifted to the surface with a block and tackle. Lifting takes place very slowly for the first 0.5 m, in order to avoid suction at the base of the sample. The sample is coated with layers of paraffin wax, and may be placed in a container packed with damp sawdust or other suitable material. The complete process takes about 3 h, including preparation for shipment. Tests by Lefebvre and Poulin have shown that this sampler is capable of obtaining soil of comparable quality to that produced by block sampling in the sensitive clays of eastern Canada. The sampler was used in the Bothkennar clay in Scotland, where it provided the highest quality of samples obtainable (Clayton et al. 1992). It was found, however, that the apparatus is quite time-consuming and difficult to use. Lefebvre and Poulin (1979) note that it is not intended that this technique should replace tube sampling for routine investigations. But where the highest quality samples are required for testing of soft or sensitive clays, at the time of writing this apparatus provides the best method of obtaining undisturbed samples from depth. DRIVE SAMPLERS

Drive samplers are samplers which are either pushed or driven into the soil without rotation. The volume of soil corresponding to the thickness of the sampler wall is displaced into the surrounding soil, which is either compacted or compressed. Drive samplers can be divided into two broad groups: open-drive samplers and piston drive samplers. Open-drive samplers consist of a tube which is open at its lower end, while piston drive samplers have a movable piston located within the sampler tube. Piston samplers can be pushed through a soft soil to the desired sampling level, but open-drive samplers will admit soil as soon as they are brought into contact with, for example, the bottom of a borehole. Open-drive samplers Open-drive samplers suffer from several disadvantages, as Hvorslev (1949) pointed out. Poor cleaning of the borehole before sampling, or collapse of sides of the borehole after cleaning may mean that much of the recovered soil is not only highly disturbed, but also non-representative. The use of a large area ratio can induce soil displaced by the sampler drive and causing large-scale remoulding of the

3

Undisturbed Sampling Techniques sample. The problems of pressure above the sample during the drive, and of sample retention during withdrawal have been noted in the previous chapter.

Fig. 7.2 Schematic diagram of the Sherbrooke down-hole block sampler (Lefebvre and Poulin 1979). The advantages of open-drive sampling are principally those of cheapness, ruggedness and simplicity of operation. Open-drive samplers can be arbitrarily divided into two groups. Thin-wall open-drive samplers have been defined as those with a wall thickness of sampling tube of less than 2.5% of the diameter, corresponding approximately to an area ratio of 10% (Hvorslev 1949). This classification is not a good guide to the amount of sampling disturbance because of the influence of cutting shoe taper and in situ stress level in the soil. In the following discussion thin-wall sampling devices are taken to be those with an area ratio of less than 20%, and a suitable cutting shoe taper, while thick-wall samplers are taken to have an area ratio greater than 20%.

4

Site Investigation

Fig. 7.3 The Sherbrooke sampler in operation (photograph courtesy Dr D.W. Hight). Thick-walled open-drive samplers Thick-walled open-drive samplers are widely used throughout the world. In their most common forms they consist of a solid or split sampler barrel, threaded at both ends to take a cutting shoe (typically with inside clearance) and a sampler head provided with either a check valve or vents.

BS general purpose sampler The British Standard General Purpose 100 mm Sampler (BS ‘5930:1981), commonly termed the U100 sampler, evolved during the 1930s and 1940s (Le Grand et al. 1934; Cooling and Smith 1936; Cooling 1942; Longsdon 1945; Rodin 1949). The sampler is rugged, cheap and will provide a core sample in most British clays, which are typically heavily overconsolidated. Its size and form were adopted because of the common borehole size at the time of its development, and because of its ability to sample (however ‘inadequately) in stoney very stiff glacial clays. The British Standard U100 sampler will fit inside a 150mm dia. borehole. Harding (1949) noted that: some believe that the smaller the [casing] tube the cheaper will be the hole. This is a fallacy. In British gravel-laden deposits, nothing less than 6-inches [152 mm] diameter is worthwhile. This permits the average type of stone to be brought up by shell without pounding with a chisel and also allows of 4-inch diameter sampling.

Light percussion drillers will often use 200mm tools in preference to the 150mm size because of their greater weight, and thus their improved ability to make fast progress.

5

Undisturbed Sampling Techniques

Figure 6.11 shows a typical British 100mm dia. open-drive sampler. It has a 104 mm inside dia. at the base of the cutting shoe, a 27% area ratio, and an inside clearance of 1.4% provided by tapering the inside diameter of the cutting shoe at an angle of 30 to meet the 106mm internal diameter of the sample tube, or by providing a uniform internal diameter to the cutting shoe and thus stepping out, abruptly at the junction of the shoe and sampler tube. The outside cutting edge taper may be 20° up to a thickness of 2.3 mm, and 7° thereafter, or alternatively may initially be 30°, and then 15° up to a 6.5mm thickness. Designs vary according to the manufacturer, but shoes and tubes are always interchangeable. The normal sample tube length is 457 mm, but two tubes are often coupled together in order to allow debris at the bottom of the borehole to pass into the upper tube during driving. Thus the normal length to diameter ratio is about 4.5, with a possible maximum of 9. The U100 sampler head incorporates vents and a ball valve assembly to allow air or water to leave the top of the tube as soil enters at its base. The ball valve is also intended to improve the sample retention by preventing air or water re-entering the top of the tube if the sample starts to slide out. Because of the conditions under which the vents and ball valve are expected to work, it is important to ensure that the sampler head is cleaned before each sampling operation. Even though the build-up of pressure above the sample can be reduced to an acceptable level, it is doubtful if the ball-valve assembly can be effective in reducing sample losses. The so-called British Standard sampler is not as completely standardized as, for example the Swedish piston sampler. In CP 2001:1957 the maximum area ratio was specified as 25%, with an inside clearance of between 1% and 3%, and a drawing of a suitable design was given. BS 5930:1981 allows an area ratio of 30%, but other recommendations remain the same. The vent area should be not less than 600mm2 cutting shoe taper is not specified. The sampler may or may not use liners, to allow the specimen to be transported and stored in a lightweight cylinder which does not have to be able to resist driving forces. In the 1982 edition of this book we wrote that ‘most users of this type of sampler do not fit liners’. It is unfortunate that in the past decade it appears that many UK companies, apparently driven by the need to reduce costs, have taken to using plastic liners inside steel outer tubes, in order to reduce the number of metal tubes they require to hold in stock. We have repeatedly observed the severe distortions induced when plastic liners are used, in comparison with the relatively low level of distortions when they are not. Figure 7.4 shows an example of this, in a laminated clay. The inclusion of a plastic liner, typically about 3mm thck (means that the cutting shoe thickness must be increased. Taking the example of the lower cutting shoe in Fig. 6.11, it can be calculated that the area ratio will increase from 27% to 41%. Examination of cutting shoes used with this type of sampler suggests that a value of area ratio of 45—50% is usual, equivalent to a B/t ratio of about 11. As we have shown, the minimum axial strain (at the centreline) will be of the order of 4—5%, and peripheral strains and shear distortions can be expected to have a very significant effect on soil properties (Georgiannou and Right, 1994). The U100 sampler is rugged, and easy to use. The sample tubes are screw-threaded to the head and cutting shoe before sampling. The sampler head is screwed to a sliding hammer (also termed a ‘jarring link’) and lowered down to the base of the hole on square rods. The sampler is driven into the soil by repeatedly lifting the rods through about 500mm and allowing them to fall. The number of blows and the distance moved by the sampler head during the drive are recorded by the driller. The sampler may be pulled immediately from the soil, or in stiff cohesive soils it may be left in the soil for a few minutes before it is brought to the top of the hole. After sampling, the tube and soil are carefully separated from the cutting shoe and sampler head. A small quantity of soil is removed from either end of the tube if necessary, and the ends of the sample are waxed, packed and then sealed with either plastic, or screw-threaded metal caps. If damaged or blunt, the cutting shoe is replaced before the sampler is next used.

6

Site Investigation

Fig. 7.4 Heavy shear distortion in a clay sample taken using a driven BS U100 sampler with a plastic liner (courtesy Surrey Geotechnical Consultants Ltd). In many countries the use of light-weight drilling rigs, percussion equipment and large borehole diameters is not advantageous. American manufacturers of samplers provide a very wide range of sampler diameters, with a variety of lengths. The Acker Solid Tube Sampler is available in a great many sizes. Its inside diameter may be 1in. (38mm), 2in. (52mm), 2km. (64mm), 3m. (76mm), 4in. (102mm) or 5in. (127mm). Its length may be either l8in. (457mm) or 60in. (1524mm). The difference between inside and outside diameters is quoted as ½ in. (12.7 mm) and the area ratios and length to diameter ratios of the various samplers are therefore as listed in Table 7.1. Table 7.1 Open-drive sampler characteristics for a tube thickness of 0.5 in. Length/diameter ratio Internal tube Area ratio diameter (in.) (%) L=l8in. L=60in. 1½ 78 12 40 2 56 9 30 2½ 44 7.2 24 3 36 6 20 4 27 4.5 15 5 21 3.6 12 Clearly, most of these tubes have excessive area and length/diameter ratios, and will not provide undisturbed soil for laboratory testing to give soil properties relevant to in situ conditions. However, in soil conditions where small diameter boreholes can be very much more economical than holes of 150mm dia. and larger, they are particularly useful. Terzaghi (1939) indicated the importance of what he described as the ‘variation survey’, a completely sampled profile of soil along several vertical lines of a site: In 1925, under the illusion that soil strata really are fairly homogeneous. I had the habit of requesting

7

Undisturbed Sampling Techniques one ‘dry [i.e., tube] sample’ for every 5 or 10 feet of test borings through homogeneous strata. Since that time the painstaking investigations of Mr A. Casagrande have destroyed my cherished illusion. No longer is there any doubt that homogeneous beds of clay are very rare Due to the universal absence of homogeneity the essential prerequisite for selecting representative samples consists in securing complete data on the variation of at least one property of the soil along several vertical lines. The samples for the more elaborate soil tests are then selected in such a manner that the properties of the most frequent soil types are determined. The weighted average of the test results is obtained by statistical methods.

The variation survey was originally accomplished by obtaining continuous 2 in. (50 mm) open-drive samples with a thin-wall sampler 30in. (450mm) long. The samples were cut into 6 in. sections before being tested for moisture content or compressive strength. In the UK, despite the advice from Terzaghi and from others (for example, Rowe (1968, 1972)), few site investigations involve taking continuous samples of any sort. Presumably this is because of a desire to reduce the cost of site investigation. Such an attitude must be considered very short-sighted. Both of the sampler types discussed above have ‘solid’ tubes; their sampler barrels are continuous around the circumference. Thick-walled split barrel samplers A further common sampler is the thick-walled open-drive split barrel sampler. Here the sampler barrel is split longitudinally into two halves. During driving these are held together by the shoe and head which are screwed on to each end. The split barrel allows easy examination and extraction of the sample, but makes the sampler considerably weaker. To compensate for this, such samplers are usually short, and have a high area ratio. One of the most common thick-walled open-drive split barrel samplers is used during the Standard Penetration Test (see Chapter 9). During this test the sampler is driven into the soil by repeated blows of a 65 kg hammer falling freely through 760 mm, and the number of blows required to drive the sampler a distance of 300mm is recorded as the SPT N value. The N value is assumed to be dependent on relative density in granular soils, and undrained shear strength in cohesive soils. Figure 7.5 shows the apparatus used in the UK for the SPT test. The dimensions of the sampler are defined in BS 1377:1975. Any sample obtained from this sampler will be highly disturbed, because the SPT split barrel sampler has an area ratio of about 100% and a length to diameter ratio of 13. No inside clearance is used. Samples of fine soils obtained from the apparatus should be considered as remoulded. Samples f coarse granular soils must be considered unrepresentative, because the coarser particles will not be able to enter the barrel during driving. For this reason the Report of the Subcommittee on the Penetration Test for use in Europe allows the sampler cutting shoe to be replaced by a solid steel cone with a J0 apex angle, when drilling in gravelly soils. This avoids cutting shoe damage, and prevents high penetration resistances due to the lodgement of large particles in the end of the shoe. Because of its design, the SPT thick-wall open-drive split barrel sampler will give low recoveries in most soils. It is therefore unsuitable for obtaining continuous representative samples for a variation or reconnaissance survey. In the UK the resulting sample is normally broken into lengths of about 75 mm and placed in a small disturbed sample container such as a glass jar. Samplers using liners inside the sample tube were termed ‘composite samplers’ by Hvorslev (1949). Liners allow considerable savings to be made because the structural outer sampler barrel, which transmits the driving force to the cutting shoe, can be used repeatedly. Only the liner is removed, 8

Site Investigation complete with sample, and sealed and transported to the laboratory. Where cohesionless or very soft soils are to be sampled it is, of course, necessary that they are not removed from the sampler tube before they are to be examined. We have already noted (above) the increase in the use of plastic liners in the BS U100 sampler in the UK during the last decade. In other countries better-designed liners are typically made of metal.

Fig. 7.5 Standard Penetration Test’ equipment. The Acker-split tube sampler is available with either solid or sectional liners. Sectional liners can be very useful in reducing the need for a laboratory extruder to remove soil from the tube, and they also allow the soil to be examined in the field, if necessary. If an extruder is not required for sample extraction, the liners can be used successfully in a wider range of soils. The inside diameter of the cutting shoe is 2 15/32 in. (62.7mm), while its outside diameter is 3 ¼ in. (82.6mm), giving an area ratio of 74%. The inside diameter of the liner is 2 ½ in. (63.5mm), giving an inside clearance of 1.6%. Thin-walled open-drive samplers The thin-walled open-drive sampler, or ‘Shelby Tubing’ sampler was introduced in the USA in the late 1930s (Terzaghi 1939; Hvorslev 1940, 1949). ‘Shelby Tubing’ is a trade name for hard-drawn, seamless steel tube manufactured by the National Tube Company of the USA. Early devices took two forms. The US Engineer Office, Boston District, sampler (Fig. 7.6) attached the thin-wall sampler tube to the head by spot welding it to a short length of heavy tube, which in turn threaded into the head. Another method of fixing the tube to the head, suggested by H. A. Mohr. uses tubing which is a close fit over the lower section of the sampler head, and which is fixed to the head by two Allen set screws which, when engaged, lie flush with the outer surface of the sampler tube. This design has been

9

Undisturbed Sampling Techniques incorporated into a large number of samplers, and is now in use worldwide (see, for example ASTM D 1587—74).

Fig. 7.6 Two early thin-walled open-drive samplers (from Hvorslev 1940). Early thin-walled samplers, such as those in Fig. 7.6 had relatively small area ratios (approximately 10—14%), but had length to diameter ratios of 15—20 and did not use inside clearance. The cutting edge was either cut square, or bevelled. Most modern thin-walled tubes are drawn in, in order to provide a suitable inside clearance (Fig. 7.7) and are sharpened.

Fig. 7.7 Typical detail of thin-walled open-drive sampler, showing drawn in, sharpened cutting edge. Thin-walled open-drive sample tubes are readily damaged, either by buckling or blunting or tearing the cutting edge, when they are driven into very stiff, hard, or stoney soils. Pushing, rather than hammering, tends to reduce the chances of damaging the tube. When the cutting edge is damaged, the tube must be sent to the metal workshop for reforming. In the UK, thin-walled open-drive samplers have, during the past decade, become used for the high

10

Site Investigation quality tube sampling of very stiff and hard clays, such as the London clay. Such a sampler, which is hydraulically jacked (from a frame at ground surface) into the bottom of the hole has been described by Harrison (1991). They are in wide use elsewhere, and can, with a certain amount of care, be used to obtain undisturbed samples from very soft soils with undrained shear strengths of the order of 5 kN/rn2. In very soft sensitive soils sampling will normally need to be carried out with a piston sampler.

Thin-walled open-drive sample tubes are typically 24—3Oin. (i.e. 610—762mm) long, and give a maximum sample length 2—3 in. (52—70 mm) less than this. They may be expected to have an inside clearance of up to 1— 1%. Available in a wide variety of diameters, they typically have area ratios similar to those in Table 7.2. Table 7.2 Thin-walled open-drive sample tubes Internal tube Area ratio Length/diameter diameter (mm) (%) ratio 48 15 11.5 60 13 9.1 73 12 7.5 86 10 6.4 121 8 4.9 It can be seen that for this type of sampler, the 121 mm internal diameter tube provides an excellent combination of low area ratio and low length to diameter ratio, which would give acceptable results with a minimum of inside clearance. At the other end of the scale, the small diameter sampler is very similar to that used by Terzaghi for his variation surveys, with the improvements of a sharpened cutting edge and inside clearance. Laval sampler Probably the most effective tube sampler available for sampling soft and sensitive clays is the Laval sampler (La Rochelle et al. 1981). The use of such an expensive, time-consuming and delicate sampling process for routine sampling probably cannot be justified, but it has been shown (not only by the originators, but also as a result of trials at Bothkennar in Scotland, reported by Clayton et al. (1992)) that this sampler recovers soft and sensitive soil almost of the quality that can be achieved using block sampling techniques. The Laval sampler is shown in Figs 7.8 and 7.9. The device consists of a thin-walled sampling tube mounted on a sampler head, and housed within an external corebarrel. The sampling tube contains a screw-type head valve which ensures that an effective vacuum can be achieved above the sample during withdrawal from the ground. The external corebarrel is used to remove soil around the sampling tube, after tube penetration, to ensure that no vacuum exists at the bottom of the cutting edge during sample withdrawal. No inside clearance is required because, in the soil types in which it is intended to be used, the shearing action between the tube and the soil leads to positive excess pore pressures, a reduction locally in effective stress, and a consequent lubricating effect. The inclusion of inside clearance was thought by La Rochelle et al. to introduce unnecessary ‘squeezing in’ of additional soil, and consequent disturbance. The sampling tube is precision machined from ZW-1035 carbon steel tubing with an i.d. of 200mm and an o.d. of 218 mm, to give a uniform circular internal cross-section along its length, and an internal diameter of 208 ± 0.03 mm. With a wall thickness of 5mm, the area ratio is 10% and the B/t ratio is 42. The cutting edge angle is 5°.

11

Undisturbed Sampling Techniques

Fig. 7.8 The Laval sampler (photograph courtesy Dr D.W. Hight). The operation of the Laval sampler is shown in Fig. 7.10. A borehole is made to the required depth, either open-hole using a fishtail bit, or as a result of previous sampling. No casing need normally be used, since bentonite mud flush provides wall support. The sampling assembly is lowered to the bottom of the hole with the sampler hooked on to the collar inside the top of the corebarrel (Fig. 7.10). With the head valve open, the sampling tube is gently unhooked by lifting and turning the inner rod at the surface. The sampler is pushed slowly into the soil, stopping some 50mm before contact is made between the bottom of the sampler head and the upper surface of the soil (Fig. 7.l0b). The head valve is then closed, and the corebarrel is used in conjunction with mud flush to clear soil from around the outside of the sampler tube, and to a depth of approximately 20mm below the bottom of its cutting edge (Fig. 7.l0c). The sample tube is rotated through about 90° in order to shear the soil at its base, and is then pulled gently out of the soil and hooked back on to the internal collar (Fig. 7.9). The assembly is finally removed from the borehole. The sample is extruded immediately, and is cut into slices 130mm or 200mm high which are placed on waxed plywood board and sealed in several layers of Saran paper sandwiched between brushed paraffin wax/vaseline mixture. The cost of the steel sampling tubes is such that they cannot economically be used for sample storage.

Fig. 7.10 General operation of the Laval sampler (La Rochelle et al. 1981).

12

Site Investigation

Fig. 7.9 The Laval 200mm diameter tube sampler (La Rochelle, Sarrailh, Tavenas, Roy and Leroueil 1981).

Piston drive samplers Piston samplers were developed in both Europe and the USA in the period between 1900 and 1940. All piston samplers have a piston contained within the sample tube, which is moved upwards relative to the sample tube at some stage of the sampling process. Because there may be various reasons for including this piston in the design, however, the mechanisms associated with the piston movement are numerous. Pistons have been included in sampler designs in order: 1. to prevent soil entering the sampler tube before the sampling position is reached. Many piston samplers have been specifically designed to be pushed, without a pre-drilled borehole, through the soil to the desired sampling depth (Bastin and Davis 1909; Olsson 1925, 1936; Petterson 1933; Porter 1936, 1937, 1939; Stokstad 1939), although it should be remembered that when this is done the upper part of the sample will normally be highly remoulded.

13

Undisturbed Sampling Techniques 2. to reduce losses of samples, by providing an efficient airtight seal to the top of the soil in the tube during withdrawal. Any tendency of the sample to slide out of the tube is counteracted by pressure decrease above the sample, (for example, see Ehrenberg (1933)). 3. to reduce the entry of excess soil into the tube during the early stages of sampling, as a result of using a relatively high area ratio, and to prevent too little soil entering the sampler at the end of the drive, as a result of the build-up of internal friction; and 4. to increase the acceptable length to diameter ratio. Adhesion between the tube and the soil entering it will tend to reduce recovery once large length/diameter ratios are reached, but the movement of the top of the sample away from the underside of the piston will form a vacuum which will tend to increase the recovery. Early American piston samplers, such as the Davis peat sampler (Bastin and Davis 1909), differed from early Swedish piston samplers (for example, Olsson (1925)) because the piston in the former was retracted to the top of the sample tube before it was driven, while in the latter case the piston remained fixed at the same level relative to ground surface throughout the drive (Hvorslev 1940). Hvorslev (1949) defined three main groups of piston sampler: (i) free piston samplers; (ii) retracted piston samplers; and (iii) fixed piston samplers. Free piston samplers Free piston samplers have an internal piston which may be clamped during withdrawal of the sampler, and during driving of the sampler to the required sampling depth. However, when the sample tube is being pushed into the soil during sampling the piston is free to move both with respect to the sample tube and to ground level. Figure 7.11 shows the Ehrenberg piston sampler and the Meijn piston sampler. In the former, the piston is free to move upwards at all times, but it cannot move downwards relative to the sample tube. It is not suitable for pushing through soft soil in order to reach the desired sampling level. In contrast, the Meijn sampler holds the piston at the bottom of the sample tube with two lugs on the inner rod which locate below a grooved collar in the sampler head. When the sampler reaches the correct level the sampler head and tube are rotated through 90° and the lugs clear the collar. After pushing the sample tube into the soil, loss of material is prevented during withdrawal by a cone clamp in the sampler head which prevents the piston rod sliding downwards through the head.

Fig. 7.11 Two types of free piston sampler (Ehrenberg 1933; Huizinga 1944).

14

Site Investigation

Free piston samplers overcome most of the disadvantages of the open-drive type of sampler, but they remain easy to use. Their main advantages are that they can be designed so that they can be pushed through debris at the base of a borehole, and that sample losses are greatly decreased by the provision of an efficient seal at the top of the sample. Despite this they are not used in the UK, perhaps because of fears that friction between the piston packing and the inside of the sample tube may cause sample compression or significantly reduce recovery. Retracted piston samplers Retracted piston samplers use the piston primarily to prevent the entrance of unwanted soil during the process of pushing the sampler to the required sampling depth (Bastin and Davis 1909; Porter 1936, 1937, 1939; Stokstad 1939). Once this depth is reached the piston is retracted to the top of the tube, and the sampler is then driven into the soil. The retraction of the piston may cause soft soil to flow upwards into the tube, and during driving a large area ratio may lead to the entry of excess soil into the tube. This type of sampler is not in use in the UK. It retains several of the disadvantages of the opendrive sampler, and is more difficult to use. Fixed piston samplers Fixed piston samplers can be used with or without a borehole. The sampler is pushed to the level at which sampling is to start with the piston rod fixed relative to the sampler head and tube, and located at the base of the tube to prevent the entry of soil. At this point the piston is freed at the sampler head, but refixed at the ground surface to the drilling rig or to a suitable frame in order to prevent it moving vertically during sampling. The sample tube is then pushed ahead of the piston into the soil. After sample driving, the inner rods extending to the ground surface from above the sampler head can be removed, since the piston is prevented from moving downwards relative to the sample tube by a clamp located in the sampler head. Fixed piston samplers have all of the advantages discussed above: they prevent the entry of debris before sampling, they reduce the entry of excess soil during sampling and they largely eliminate sample losses. Hvorslev (1949) commented that ‘the drive sampler with a stationary piston has more advantages and comes closer to fulfilling the requirements for an all-purpose sampler than any other type’. Its disadvantages lie principally with its cost and complexity in use. The original fixed piston sampler developed by Swedish engineer, John Olsson (Olsson 1925, 1936) controlled the movement of the piston before and during sampling by extending the piston rod to the ground surface inside the outer rods, and clamping it either to the outer rods or a frame at ground level. Modifications and improvements were tried over a period of many years (Petterson 1933; Bretting 1936; Kjellman 1938; Fahlquist 1941; Hvorslev 1949; Osterberg 1952; Hong 1961). In 1961, the Swedish Committee on Piston Sampling produced their findings (Swedish Geotechnical institute Report No. 19) and Kallstenius gave precise details of the apparatus in use. Since that time, work has been carried out to assess the effectiveness of piston samplers in obtaining good quality undisturbed samples (Berre et al. 1969; Schjetne 1971; Holm and Holtz 1977) and modifications to the apparatus have continued (for example Osterberg (1973) and Tornaghi and Cestari (1977)). The Japanese Society for Soil Mechanics and Foundation Engineering have subsequently published a Draft Standard for stationary piston sampling (Mon 1977). Piston samplers with fixed pistons are available with a variety of sampler barrels. These may be thinwalled (made of either seamless steel tubing or of aluminium tube) or of the composite type. Figure 7.12 shows a thin-walled seamless steel tube type fixed piston sampler similar to those described by Hvorslev (1949). The sampling tube has a rolled and reamed cutting edge. The sampler may be pushed through soft soils to the desired sampling level and, during this process, the conical piston is held at the base of the sampler tube. This is achieved by attaching the piston rod to the upper

15

Undisturbed Sampling Techniques part of the head via a few turns of the left-hand thread of the piston rod screw clamp. When the sampling level is reached, the piston rods are turned clockwise at ground surface, tightening the rods above the sampler but disengaging them at the screw clamp. The piston rod is then fixed to the rig or a frame at ground level, and the hollow outer rods are pushed smoothly downwards to drive the sampling tube ahead of the piston into the soil. After sampling, the piston rods can be unclamped and the sampler pulled to the surface using the outer rods. The piston is held up by the ball cone clamp in the sampler head. Once the sampler is at ground surface, the tube is released from the head by screwing the Allen set screws inwards. The ball cone clamp must be released by turning a screw on the side of the head through 900 before the sample tube can be pulled from the head. The vacuum release screw must be slackened before the piston can be pulled out of the sample tube. The various screws must be reset before the next sample is taken.

Fig. 7.12 Thin-walled seamless steel tube fixed piston sampler. A device such as that described above has been widely and successfully used in very soft and sensitive clays in Norway. The Norwegian Geotechnical Institute 54mm dia. sampler has the following characteristics (Void 1956; Berre et al. 1969): maximum tube length maximum sample length outside diameter area ratio inside clearance maximum length/diameter ratio

880 mm 725 mm 57mm 11—12% 1.0—1.3% 13.4

16

Site Investigation outside cutting edge angle

120

The NGI sampler has also been built in a 95mm version, giving 1000mm long samples (Berre et al. 1969). Evidence from oedometer tests suggests that 50cm2 specimens from 95 mm dia. piston samples give much less scattered test results than specimens from 54mm piston samples, in soft to firm clay. This may be a consequence of small-scale heterogeneity, or of reduced sampling or extrusion disturbance. An adaptor kit is available from the UK company, Engineering Laboratory Equipment Ltd, to allow the 54mm NGI/Geonor sampler to take 101mm dia. sample tubes with lengths of either 457mm or 1000mm. These tubes are aluminium, with an outside bevel and no inside clearance. They have the following characteristics: maximum sample length outside diameter area ratio length/diameter ratio

330 or 875mm 105mm 8% 3.3 or 8.8.

These types of piston sampler have been used widely and reasonably successfully in the UK, where typically they are pushed into the soil at the base of a borehole. Their major disadvantage lies in the slow speed with which they can be used. Inner and outer rods are screwed on in I m lengths, and the complex holding mechanisms require careful use. To speed the sampling process several researchers have proposed mechanisms which allow a fixed piston sampler to be used with only one set of rods. Bretting (1936) developed a sampler with an integral hydraulically actuated piston, intended for use in a cased borehole. The piston was fixed to the outer barrel of the sampler and held at the required level by drill rods extending to the ground surface. The sample tube extended over the fixed piston to a second (upper) piston, which could be forced downwards by the application of water pressure to the inside of the drill rods at ground level. This principle was subsequently also used by Osterberg (1952, 1973) for his hydraulic piston sampler (see Fig. 7.13).

Fig. 7.13 The Osterberg composite hydraulic fixed piston sampler (Osterberg 1973).

17

Undisturbed Sampling Techniques

The original Osterberg thin-walled hydraulic piston sampler was available in 127mm dia. (6% area ratio) and 72mm dia. (6.3% area ratio) forms. Raymond et al. (1971) and Raymond (1977) used this sampler with a restricted sample length of 600mm in the sensitive Leda clay and found it to give results second only to block sampling. The improved Osterberg composite hydraulic fixed piston sampler (Osterberg 1973) gives a maximum sample length of 1625mm with a diameter of 127 mm. It has an area ratio of 18%, inside clearance of 0.4%, a length to diameter ratio of 12.8 and a cutting edge angle of 7°. It has vents in the sample tube to allow vacuum relief below the tube when pulled from the soil, and the sampler can be rotated before pulling, in order to shear the soil at the base of the tube. Fixed piston samplers are also widely used in Europe and the USA with liners. These composite fixed piston samplers have the advantages that they can be made more rugged and therefore more suitable for displacement boring, but the tube used to retain and store the soil after sampling can still remain lightweight and cheap. Evidence from sampling the soft, sensitive Leda clay (Raymond et al. 1971) suggests that the use of liners may be essential in small diameter samplers in some soil types, since both extrusion forces and the vibrations caused by sawing seamless steel tubing cause severe disturbance. The Swedish composite fixed piston sampler (Kallstenius 1961) has a 700 mm stroke and a 50mm dia. at the cutting edge. Its outside diameter is 60mm, with plastic sectional liners of 170mm length and 50.2 mm inside diameter used to retain the sample. The base of the piston is 50° cone ended. Its characteristics are therefore: maximum sample length internal diameter at cutting shoe area ratio inside clearance length to diameter ratio

700mm 50mm 44% 0.4% 14

Swedish experience has shown that very high quality samples can be obtained even with such a high area ratio and length to diameter ratio. Kallstenius (1958) showed the importance of cutting shoe design: for the Swedish standard piston sampler the taper is specified as 45° up to a thickness of 0.3mm and 5° thereafter. Similarly, in the USA, the Lowe—Acker composite fixed piston sampler (Lowe 1960) eliminates the effects of the high area ratio of a composite barrel by coupling a 150mm length of thin-walled tube to a 75mm long tapered coupling at its base. Despite a 70mm sample diameter and an area ratio of about 60%, Lowe (1960) has claimed that the effect of the thicker walled barrel section of the sampler is negligible. The Swedish standard piston sampler is of rather small diameter compared with many modern devices, and the importance of a large diameter specimen in reducing sample disturbance and obtaining representative samples has already been noted. Holm and Holtz (1977) have presented the results of a study in which the Swedish Standard 50mm dia. device was compared with the NGI 95mm sampler, the Osterberg 127mm hydraulic piston sampler, and a 124mm dia. research sampler developed at the Swedish Geotechnical Institute. The results of this study indicate no significant differences between either the ratio (preconsolidation pressure/in situ vertical stress) or undrained shear strength derived from laboratory tests on specimens obtained by the various devices, but there are indications that: 1. results of oedometer tests on 50mm samples are more scattered, supporting the findings of Berre et al. (1969); and 2. the undrained modulus obtained from 50mm samples may be significantly lower. Foil and stockinette samplers Several devices have been developed to allow very long samples to be taken. Long samples are

18

Site Investigation particularly desirable when soil is highly variable, containing for example interbedded clays and sands. Begemann (1974) has argued that in these conditions the calculation of settlement or predictions of the effectiveness of vertical sand drains cannot be made without a full, undisturbed, detailed picture of the soil, and this view is certainly supported by the work of Rowe (1968a, b, 1972). Inside friction can be reduced by the cautious use of inside clearance; it cannot be eliminated without causing serious disturbance to the soil inside the sampler as a consequence of the reduced support. The provision of sliding liners within the sampler barrel means that lateral restraint can be maintained while frictional forces between the soil and its container are eliminated. Two types of device are available: that developed by the Swedish Geotechnical Institute and described by Kjellman et al. (1950) inserts aluminium foils between soil and sampler, whilst a sampler developed at the Delft Soil Mechanics Laboratory by Begemann (1961, 1971) surrounds the soil with a nylon stocking reinforced plastic skin and supports it with bentonite fluid. The principles of operation of the Swedish foil sampler are shown in Fig. 7.14. According to Broms and Hallen (1971), two types of sampler exist, giving sample diameters of either 68mm or 40mm. There are sixteen rolls of very thin high strength steel foil in the sampler head of the 68mm sampler. Each foil is 12.5mm wide, and about 0.1mm thick. The thickness may be varied depending on the required maximum sample length and the anticipated frictional forces to be resisted. The sampler is pushed into the soil without a borehole. As the sampler is pushed downwards the foils, which are attached to a stationary piston, unwind from their rolls and completely surround the sample.

Fig. 7.14 Principle of operation of the Swedish foil sampler, and detail of the Mark V sampler head (Broms and Hallen 1971; Kjellman, Kallstenius and Wager 1950).

19

Undisturbed Sampling Techniques The maximum sample length that can be obtained depends on the strength of the foils and the size of the foils and their magazines. The 40mm foil sampler can hold a maximum length of 12m of foil, while the 68mm sampler can store 30m. The friction between the foils and the sampler can be reduced by lubrication when sampling clays, and it has then been found possible to obtain continuous cores more than 20m long in very soft to soft soil. In sands, lubricants may penetrate the soil and cannot therefore be used; the length of sample is reduced. The sampler is generally pushed or driven into soft cohesive soils. When silty or sandy soils are met, jetting may be needed to reduce the driving resistance. Ramming and jetting reduce the quality of the sample. Broms and Hallen (1971) describe a drilling rig for use with the foil sampler to obtain continuous samples of hard materials where rotary drilling is required. Using this equipment it has been possible to obtain cores of sand or hard boulder clay (till) up to l0m long. The 66mm dia. Delft continuous soil sampler is shown in Fig. 7.15. In the early 30mm dia. version of this sampler (Begemann 1961, 1971) the soil entering the tube was primarily prevented from collapsing by bentonite fluid pressure. A stocking stored on the ‘stocking tube’ was surrounded by red vulcanizing fluid held in the chamber between the stocking tube and the outer sampler tube. The end of the stocking was secured to a cone-ended piston fixed at ground level. The inside of the sampler was filled with bentonite—water slurry. As the sampler was pushed into the ground the stocking unrolled from the stocking tube and rolled on to the outside of the soil. Contact between the slurry and the red vulcanizing fluid caused it to solidify, thus making a water-tight container, and preventing lateral strain of the soil.

Fig. 7.15 66mm dia. Delft stocking sampler (Begemann 1974). The sampler was originally pushed into the ground with a 10 tonne penetrometer, such as is used for the cone penetration test. The maximum sample length was determined by the maximum penetration

20

Site Investigation force available, and the length of stocking that could be stored. This varied between l0m and 20m. The sampler was advanced in 1 m drives, and after each section of tube was added it was filled with bentonite slurry. When the desired depth was reached, the bottom of the sample was closed by a diabolo valve operated by rotating the tubes at ground surface, to prevent loss of bentonite and soil during withdrawal. The 66mm diameter continuous sampler (known in The Netherlands as a Begemann boring) provides samples large enough for laboratory consolidation and triaxial testing (Fig. 7.16). A bentonite flush, weighted with barytes to increase its density, is used in order to provide support for the sample as it passes the opening through which the stocking emerges. In soft soils the density apparently may need to be modified so that the imposed horizontal stress does not exceed that in the soil. The larger sampler includes a plastic liner which virtually eliminates the space between the soil and the inside of the extension tubes. These liners also provide support as the samples are transported to the laboratory. The maximum sample length for the 66mm sampler is 19 m according to Begemann (1974), and a 17 tonne penetrometer has been used.

Fig. 7.16 Continuous l0m long sample in soft soil obtained using the Delft sampler (photograph courtesy Delft Geotechnics Laboratory). Foil and stockinette samplers are relatively expensive, and are best reserved for soft, primarily cohesive soil types. They are no widely used in the UK, perhaps because large thicknesses of soft, fine sediment are not particularly common. Whilst they are very useful in providing a continuous record of

21

Undisturbed Sampling Techniques complex soil conditions, they will not normally be expected to give as high quality samples as may be obtained using the best methods now available. ROTARY SAMPLERS

Chapter 5 described the general principles of rotary coring, as carried out routinely during site investigation. The main tool used for rotary sampling of hard rocks in the UK and most of the world is the rotary corebarrel (Fig. 5.11) but rotary samplers are increasingly being adapted to sample virtually all types of soil and rock. In this section we describe the three principal types of rotary corebarrel used on stiff and hard clays, and weak, fractured and weathered rocks. Also described is a simple rotary sampler designed to give very high quality samples in soft sensitive clays. Rotary corebarrels designed to sample the harder materials encountered during site investigations can be classified into three broad groups: 1. corebarrels with retracted inner barrels, such as the conventional double-tube swivel type corebarrel (Fig. 5.11), here termed ‘retracted corebarrels’; 2. corebarrels where the inner barrel protrudes ahead of the outer barrel, in an attempt to protect the ground being sampled from the deleterious effects of flush fluid, here termed ‘protruding corebarrels’; and 3. corebarrels where the inner barrel is spring mounted, so that it protrudes in relatively soft ground, but retracts when harder layers are encountered, here termed ‘retractor barrels’. Refracted corebarrels Double-tube swivel type corebarrels British site investigation practice uses large diameter double-tube swivel type core- barrels, normally with face discharge diamond bits and a built-in core catcher. Common barrel sizes in use in the UK are NW, HWF, PWF and SWF, giving core diameters of 54.0, 76.2, 92.1, and 112.7 mm. The double-tube corebarrel contains a stationary inner barrel supported on a swivel. Flush fluid is pumped down the inside of the rods which run from the drilling rig at ground level to the top of the corebarrel. Once inside the barrel the flush fluid passes down between the inner and outer barrels and discharges through ports in the cutting face of the bit. The inner barrel is extended with a core catcher box which contains a split ring core catcher. When the barrel is pulled from the bottom of the hole, the catcher spring prevents loss of core by moving down the inside taper of the catcher box and progressively gripping the core more tightly if it slips downwards. Double-tube swivel type corebarrels of large diameter can be used with great success not only to provide good quality core of sound rock, but also to provide samples of very stiff or hard clays. Once the core enters the inner barrel it is protected from erosion of the flush water and from the torsional effects of rotation. The top of the inner barrel is vented to prevent build-up of pressure over the top of the core. Should this vent become blocked, the pressure in the inner barrel may prevent core entry after as little as 0.5 m of coring, and in softer formations the core may be washed away or ground away. Wireline drilling techniques, coupled with polymer mud, are now frequently used in the stiff and hard Eocene formations of the London area (the London clay, the Woolwich and Reading beds, and the Thanet sand), as an alternative to thin-wall tube sampling, when higher quality samples than can be obtained using the BS U100 tube sampler are required. Corebarrels tend to have a larger area ratio and inside clearance than is generally accepted for drive samplers. The former is an advantage, because one of the problems when drilling in soft formations is 22

Site Investigation to keep the pressure between the bit and rock or soil low enough to prevent the barrel fracturing or displacing the material beneath it. The larger inside clearance, which might be 2.4% for an HWF barrel, can cause serious problems. Although the core is protected from erosion by the flush fluid once it enters the inner barrel it is still in contact with that fluid and shales, mudstones and clays may deteriorate significantly if water flush is in use. Because the core is not well supported in the inner barrel, the effects of vibrations will be severe. It is clear that when the highest possible recovery is required, the normally accepted rules to obtain fast economical progress with a diamond drill cannot be followed. Low bit pressures may reduce bit life by polishing the diamonds, and the low rotational speeds necessary to prevent vibrations from damaging the core will reduce the penetration speed. The main disadvantage of the double-tube swivel type corebarrel is that considerable skill and experience are required to use it successfully. When soil conditions are difficult both equipment and technique must be chosen with care: flush fluid, rig stroke, barrel length, diameter and design, and bit type will all be important. The correct flush fluid can slow or even prevent the disintegration of the core. A long- stroke rig helps to reduce erosion and softening of the core by reducing the need for rechucking of the drill rods, and also lessens the chances of blocking if the flush pump is stopped during rechucking. Short barrels (say 1.0— 1.5 m long) of large diameter may be preferable to long thin barrels, because the effects of vibration and softening will be reduced. The use of liners of either rigid plastic or flexible plastic (Mylar) helps to reduce the effects of flush fluid, and largely eliminates damage to the core during withdrawal from the inner barrel. They can reduce inside clearance, and their smooth surface allows easy entry of the core. Much damage can be done to good core during its extraction from the barrel. It is not uncommon to see core removed by holding the corebarrel almost vertical on a wire rope, and repeatedly hitting the inner barrel with a hammer. This method is not only likely to damage the inner barrel, but also will often damage the core. There is little or no chance of the driller being able to maintain the pieces of core in the same relative orientation as they occupy in the barrel while he struggles to place them in a corebox. Core should be extruded with the corebarrel held horizontally, using a coreplug in the inner barrel. Pressure should be smoothly applied to the back of the coreplug so that the core is extended with a minimum of vibration into a plastic receiving channel of about the same diameter as the core. After extrusion both core and plastic channel should be wrapped in clear polythene sheet and securely taped before being placed in the corebox. Iwasaki et al. (1977) and Seko and Tobe (1977) have carried out comparative trials between doubletube swivel type corebarrels and a variety of other sampling devices. Iwasaki et al. found that a double-tube swivel type corebarrel with a face discharge bit, modified with a built-in check valve rather than a spring core catcher, an inner tube stabilizer and a reduced inside clearance of about 1.4% would give better results than a Denison sampler (see later) for clays with undrained shear strengths in excess of 150 to 200 kN/m2. Seko and Tobe carried out comparative sampling trials in the very stiff and sometimes hard clays of the Tokyo area: they used the samplers listed in Table 7.2. Table 7.2 Samplers used in comparative sampling trials Diameter of core Type of sampler (mm) 1 Double-tube swivel type corebarrel with a tungsten bit 60 2 Denison sampler 3 Retractor barrel, without a core catcher 70-80 4 Retractor barrel, with core catcher 5 Wireline Denison sampler 6 Single-tube corebarrel 7 Thin-wall hammered open-drive sampler

23

Undisturbed Sampling Techniques All the rotary samplers were used with mudflush. It was found that the double-tube swivel type corebarrel gave the best specimens, based on unconfined compressive strengths and modulus of elasticity values. Types 2, 4 and 5 were classed as second best, but thought unreliable, often causing serious disturbance. Surprisingly, the retractor barrel without the core catcher gave worse results than that with a spring core catcher. Significantly, the single-tube corebarrel and the hammered thin-walled open-drive sampler were described as ‘entirely unsuitable’. Protruding corebarrels In order to reduce the effects of flush fluid and torsional forces on the core, a number of devices have been developed in which the inner barrel extends below the bottom of the rotating corebit. The first of these devices was developed by the Ministry of Railways in Japan in the mid-1930s (Hvorslev 1940; Iwasaki et al. 1977). Subsequently Johnson (1940) reported the development of a similar corebarrel with a protruding inner barrel for taking samples of dense but erodible soils in the Denison District, Texas. Denison corebarrel The Denison corebarrel (Fig. 7.17) is a triple-tube swivel type corebarrel, with a shoe with a sharp cutting edge threaded onto the inner barrel and extending below the cutting teeth of a tungsten corebit. The length of the corebit must be changed to alter the amount by which the shoe extends below the corebit. According to Hvorlev (1949) and Lowe (1960) a 50—75 mm inner barrel protrusion is suitable for relatively loose or soft soils, whilst the cutting edge should be flush with the corebit in ‘very stiff, dense and brittle’ soils. The Denison corebarrel uses a ‘basket’ type spring core catcher, where a number of curved, thin, flexible springs are fixed to a base ring by rivets, or by welding. According to Hvorslev (1949) the use of such thin springs means that the core catcher is frequently damaged and must be replaced. The inner barrel encloses a liner, often of brass. The original inner barrel design by Johnson (1940) had a 32% area ratio, and a 0.6% inside clearance. The use of a high area ratio means that samples of hard clays and dense sands or gravels will be greatly disturbed, and better sampled by a conventional retracted inner barrel type sampler. Very soft to firm clays can be more effectively sampled with a fixed piston sampler. According to Lowe (1960) the Denison sampler is designed for use in stiff to hard cohesive soils and in sands. It is rarely used in the UK, where stiff clays are sampled using the 100mm thick-walled open-drive sampler, and the undisturbed sampling of sands is rarely attempted.

Fig. 7.17 Denison triple-tube corebarrel (Johnson 1940).

24

Site Investigation

Retractor barrels One of the problems facing the Denison corebarrel user is that the inner barrel protrusion must be preselected. To overcome this problem, several corebarrels have been developed which include spring mounted inner-barrels. Pitcher sampler The pitcher sampler (Terzaghi and Peck 1967; Morgenstern and Thomson 1971) is shown in Fig. 7.18. The inner barrel consists of a thin-walled sampler tube with a rolled and reamed cutting edge which is fixed to the inner head by set screws. The outer barrel has a tungsten insert corebit. The inner head is not fixed to the outer barrel; when the device is lowered to the bottom of the hole the head is supported immediately above the bit and flush fluid can be passed down the drill rods through the centre of the sample tube to remove any debris left at the bottom of the hole. Once the sample tube beds on to the soil at the bottom of the hole, the central tube on the top of the inner head mates with the outer barrel head. Flush fluid is now routed via the outside of the sampling tube, and the space above the sample is vented via the top of the sampler. The lead of the tube cutting edge is governed by the spring stiffness and the hardness of the soil.

Fig. 7.18 Principal features of the Pitcher sampler.

25

Undisturbed Sampling Techniques

In theory this type of sampler is ideally suited for drilling in soils with alternate hard and soft layers. In practice it has been found that hard friable soils, such as the weathered Keuper marl, can be sampled very successfully but frequently damage the rather light sampling tube. Thus this device is likely to be unsuitable for sampling inter-layered soil and rock, for example inter-layered clay and limestone. More rugged forms of retractor corebarrels have been developed in Australia by Triefus, and in France by Soletanche (Cambefort and Mazier 1961; Mazier 1974). Mazier corebarrel The Mazier corebarrel (Fig. 7.19) is a triple-tube swivel type retractor barrel, whose effectiveness (as with the Pitcher sampler) relies on the fact that the amount of inner barrel protrusion is controlled by a spring placed in the upper part of the device. The inner barrel contains a brass liner which can be used to transport samples to the laboratory, or for storage. The cutting shoe on the bottom of the inner barrel is substantial, making it much less easily damaged than a thin-walled seamless tube, but introducing the problems of disturbance when the high area ratio shoe travels ahead of the corebit.

Fig. 7.19 Detail and principle of operation of the Mazier corebarrel.

26

Site Investigation Triefus triple-tube retractor corebarrel The Triefus triple-tube retractor corebarrel uses the same principles as the Mazier device, but the maximum projection of the cutting shoe ahead of the corebit is much smaller (between 3mm and 16mm). This means that the absolute minimum of flush fluid should be used, in order to prevent washing and scour of soft or friable formations. Unlike the Mazier barrel, no core catcher is used and the base of the barrel must be sealed by dry blocking (i.e. drilling without flush fluid) at the end of the run. The liner tubes in the Triefus barrel can be either split steel, or solid transparent plastic, the latter allowing inspection of the core in the field without removal from its tube. Additional major advantages of the Triefus barrel are that most of its components are interchangeable with those of their triple-tube standard corebarrels (allowing the standard barrel to be converted to a retractor barrel in the field), and that both these barrels are fitted with a core extractor plug (blow-out valve) to allow the core in its liner to be pushed gently from the barrel whilst in a horizontal position. Retractor barrels find their best use in formations of variable hardness, where piston drive sampling cannot penetrate and when standard rotary coring provides insufficient protection to the soil. In general, thin-walled sampler retractor barrels such as the Pitcher sampler are susceptible to tube damage, while those with thick cutting shoes are much more likely to cause serious disturbance to the soil. Where soil conditions are relatively uniform and the soil is of sufficient undrained shear strength (>150—200kN/m2) the results of Iwasaki et al. (1977) and Seko and Tobe (1977) indicate that a double-tube swivel type rotary corebarrel can provide less disturbed samples than either protruding or retracting inner tube barrels, or driven thick-walled open-drive samplers. SAND SAMPLING

‘Practical undisturbed’ samples, in Hvorslev’s terms, can be obtained from sand deposits: but despite the fact that the soil structure, water content, void ratio and constituents remain unaltered such samples are not normally suitable for compressibility testing. Indeed, Broms (1980) asserts that such samples cannot be obtained of a sufficiently high quality to allow determinations of potential liquefaction problems. The effects of total stress relief on granular soils usually result in very large reductions in effective stresses, and the properties of such soils tend to be highly dependent both on stress history and current effective stress level. But in addition, small changes in shear stress can also destroy the ‘memory’ of previous stress applications, leading to considerable reduction in stiffness, in sands. Undisturbed sand sampling can be very expensive, and is normally only required in special circumstances, for example to obtain values of in situ density for earthquake liquefaction problems or for compressibility studies. In Japan, for example, the havoc created by earthquake damage has caused a considerable interest in the sampling of granular materials and has lead to widespread research and publication (for example, Yamada and Uezawa 1969: Hanzawa and Matsuda 1977; Isihara and Silver 1977; Seko and Tobe 1977: Tohno 1977 and Yoshimi et al. 1977). Hvorslev (1949) outlined a number of techniques for sampling sand using: 1. 2. 3. 4. 5.

thin-wall fixed piston samplers in mud-filled holes; open-drive samplers under compressed air; impregnation; freezing; core catchers.

Fixed piston sampling A study of the effectiveness of thin-walled fixed piston sampling has been carried out at full-scale in 27

Undisturbed Sampling Techniques the laboratory by Marcuson et al. (1977). Using a 1.22m dia. x 1.83m high specimen of fine sand, and mud flush rotary drilling, they obtained samples using a 76.2mm dia. piston sampler with an area ratio of 11%. By comparison between soil densities when placed and after sampling it was found that samples of dense sand were slightly loosened and samples of loose sand densified, but the results suggest an accuracy within ±3.5% of the placed density for sands with relative densities of between 20% and 60%. Bearing in mind the difficulty of creating uniform samples (the authors suggest a variability of density at the time of placement of about + 3.2% to —1.6%) these results seem encouraging, and support the view of Friis (1961) regarding the value of thin-walled fixed piston sampling under drilling mud in sand. The use of compressed air to displace water from around the sampler tube in the borehole, and thus reduce the losses of samples in sand was first suggested by Glossop to Bishop (1948) and independently by Vargas to Hvorslev (1949).

Bishop’s sand sampler Bishop’s sand sampler (Bishop 1948; Nixon 1954; Serota and Jennings 1957) consists of a 63.5mm thin-walled open-drive sampler held by set screws to a head containing a rubber diaphragm check valve and vents (see Fig. 7.20). This assembly is mounted within a compressed air bell which is connected to an air pump at the ground surface. The sampler is used in the following manner.

Fig. 7.20 The Bishop compressed air sand sampler (Bishop 1948). 1. Lower the sampler inside the compressed air bell to the base of a cleaned borehole. 2. Push the sample tube ahead of the bell into the soil, using the rods. 3. Remove the rods, and force compressed air into the bell via the relief valve in the sampler head. The relief valve vents to the inside of the bell at a pressure difference of about 150 kN/m2, and this pressure bears on the upper surface of the diaphragm, ensuring that it works efficiently.

28

Site Investigation 4. Once air bubbles rising to the surface of the water in the borehole indicate that the water in the bell has been expelled, the sample tube is pulled from the soil into the bell, and then rapidly brought to ground surface using the lifting cable. 5. Remove the spacer, push the sample tube and head out of the bell, and release the sample tube set screws. 6. Cap the base of the tube and release the check valve before removing the sample tube from the head. The Bishop sand sampler uses arching and the small capillary suctions set up at the sand/air interface to reduce sample losses. This principle has since been used by Yamada and Uezawa (1969).

Soil impregnation Van Bruggen (1936) described the use of soil impregnation with a dilute emulsion of asphaltic bitumen, in order to impart cohesion to granular soils and thus allow them to be sampled. The bitumen was subsequently removed by washing through with a solution of carbon disulphide and acetone. Such a process would certainly change the properties of the soil. Hvorslev (1949) describes the use of chemical injection around the cutting shoe of an open-drive sampler to solidify soil and help to retain samples of granular material. Subsequently Karol (1970) and Borowicka (1973) have reported the use of various resin and silica grouts to prevent disturbance to the soil structure during sampling. Impregnation and injection are expensive and relatively difficult to use: the soil to be sampled cannot be impregnated unless the chemicals and so on, can be effectively removed at a later date, and in addition, most grouts and emulsions will not penetrate relatively impervious sand or silt soils. The method is therefore rarely used.

Freezing In contrast, freezing has been widely used to seal the bottom of sample tubes (once driven), to prevent disturbance to the soil during transporting to the laboratory, and to freeze soil before sampling, (for example, Fahlquist 1941; Hvorslev 1949; Ducker 1969 and Yoshimi et al. 1977). These techniques are very expensive, and yet the need for undisturbed silt and sand samples has resulted in their continued use, particularly in the earthquake areas of the Far East and the USA. A relatively economical method of obtaining frozen sand samples is shown in Fig. 7.21. A 73mm thin-wall steel tube is inserted into the ground, while removing soil from its interior with an auger. Once the desired depth of freezing is reached, the lower end of the tube is sealed with cement grout and the plastic ‘freezing tube’ inserted down the centre of the steel tube. Circulation of an ethanol and crushed dry ice coolant at -40 to -60°C results in a frozen column of soil which can be extracted by pulling the steel tube from the ground. At ambient air temperatures of 23—30°C and a ground temperature of 22°C at 0.9 m depth, Yoshimi et al. (1977) acquired columnar samples 5.8 m long and 380mm dia. after 16 h of coolant circulation.

Fig. 7.21 Method of sand sampling by freezing adopted by Yoshimi, Oh-Oka (1977).

29

Undisturbed Sampling Techniques

The value of samples obtained by freezing depends on the amount of density and soil structure change caused by the process. The amount of strain taking place during freezing increases with increasing relative density, decreasing applied pressure, and increased freezing time. Tests by Yoshimi et al. on specimens of soil placed in the laboratory and on samples obtained using the freezing tube technique described above indicate that soil adjacent to the freezing tube was significantly altered by freezing; loose sand was densified, but medium dense and dense sands loosened. Densities in the outer half of the sample, however, showed no alteration and indicated values mostly within ±2% of the average placed or field dry density. Core catchers Core catchers can be used with great success to retain granular soils, but their design may introduce considerable disturbance during the sampler drive. Spring systems such as used in rotary samplers are examples of the types of catching device which should be used with great caution. The catchers described by Yamada and Uezawa (1969) and Isihara and Silver (1977) are examples of systems which should lead to a minimum of disturbance. SAMPLER SELECTION

For reasons of economy it is sensible to adopt the cheapest sampling techniques compatible with the aims of an investigation. To do this, the reasons for drilling and sampling must be clearly defined. In many cases only very simple methods of sampling will be necessary — the priority will be the proper identification of soil type and soil fabric, and there will be little need to obtain realistic soil parameters. In other instances, for example when complex computer analyses are to be undertaken during design, much higher quality samples will be required so that sophisticated laboratory tests can be carried out. The German Standard DIN 4021 defined five quality classes of soil sample, and quality classes have also been discussed by Idel et al. (1969), and by Rowe (1972), and have been defined in BS 5930:1981. In principle they identify the range of samples shown in Table 7.3. Table 7.3 Soil sample quality classes Quality State of soil sample class No geometric distortion. Shear strength and 1 compressibility are unaffected. Geometric distortion. Density and water content 2 unaffected. Density altered. Water content and particle size 3 distribution unaffected. Water content and density altered. Particle size 4 distribution unaffected. Particle size distribution altered by loss of fines or grain 5 crushing. Based upon the research of the past decade, this classification no longer seems relevant and, indeed, it does not appear to have been used in practice. We can note, for example, that Quality class 1 samples cannot be obtained with routine tube samplers, and that the shear strength and compressibility of even the best quality block samples are likely to be affected as a result of changes in effective stress during sampling. It is difficult to envisage how water content can remain unaffected when density changes, given that most soils are likely to be saturated when in situ. And inadequate particle size distributions may result as much from a poor choice of sample size as from poor sampling technique.

30

Site Investigation In the UK the typical ‘routine’ soil sampling associated with light percussion drilling is as follows. 1. All soils. Small disturbed sample (jar sample) immediately each new stratum is detected. Water samples from every water strike. 2. Sand. Standard penetration test (SPT), with a jar sample taken from the split spoon, immediately on entering the stratum, and at 1.5 m intervals thereafter. If no sample is recovered by the split spoon a large disturbed sample (bulk bag) should be taken from the same level. Bulk bags should be taken at 1.5m intervals midway between the SPT tests. 3. Sand and gravel, or gravel. Standard penetration test with 60° cone (SPT cone), immediately on entering the stratum, and at 1.5m intervals thereafter. A large disturbed sample (bulk bag) should be taken from the level of the SPT (cone) test. 4. Chalk, marl or silt. Alternate 100mm thick-walled open-drive samples (UlOOs) and SPT tests with jar samples taken from the split spoon sampler, at 1.5 m intervals. In hard marls and cemented silts an ‘H’ size double-tube swivel type corebarrel with air flush is sometimes used. 5. Hard clay. U100 samples may be possible, but otherwise SPT tests are carried out. Exceptionally, mud flush and rotary coring is used. 6. Stiff clay. U100 sample immediately upon entering the stratum and then at 1.5 m intervals. A small disturbed sample is taken from the cutting shoe of the U100 samples, and at 1.5 m intervals midway between the U100 samples. 7. Very soft clay or peat. 54mm or 100mm dia. thin-walled fixed piston samples, either continuously in a shallow deposit, or at 1 m intervals in a deep deposit. 8. Rock. Continuous rotary core from a double-tube swivel type corebarrel, with water flush. The minimum size is usually ‘N’. This scheme leaves the ‘variation survey’, that is the assessment of soil variability, as well as many of the decisions about sampling, and the recognition of strata changes, to the drilling foreman. He is the only person to see the majority of the soil profile, and must watch the disturbed material produced by the drilling process most carefully so that subtle (but perhaps significant) features are not to be missed. This type of ground investigation, and the routine testing practices associated with it can now be viewed as barely adequate. The ground profile is not properly examined, the samples obtained are typically of low quality, and the testing carried out may often be unnecessary, or inappropriate for the engineering needs of the project. When proper design of ground investigation is carried out there is a need to match the sophistication of sampling (and in situ testing) to the sophistication of the analysis and design of the project. Empirical and semi-empirical methods of design are certainly acceptable, and should be used in conjunction with the appropriate sampling and test methods upon which they were originally based. But more fundamental methods of analysis, for example for the constitutive modelling of soil behaviour via finite element or finite difference analyses, will require high-quality sampling and testing methods. From the range of sampling methods described, the methods shown in Table 7.4 are the best available.

31

Undisturbed Sampling Techniques Table 7.4 Recommended sampling methods Recommended sampling Likely disturbance technique Very soft, soft, firm, or Laval open-drive overcored Minor destructuring sensitive clays sampler Sherbrooke down-hole block Reduction in effective stress due sampler to borehole fluid penetration Inter-layered sand, silt, Delft sampler Major loss of effective stress. clay Some destructuring Firm, stiff and very stiff Thin-walled hydraulically Minor destructuring, with clays jacked open-drive tube significant increases in effective samples stress Significant decrease in effective Very stiff and hard Wireline coring, using stress clays, mudrocks, and bentonite mud or polymer stoney clays muds with anti-swelling agents, or double-tube swivel type corebarrel with bentonite mud flush Sand Piston sampling in mud-filled Total loss of effective stress. borehole Major destructuring. Density approximately maintained Gravel Sampling from pits Only particle size and density can be obtained Weak rocks, chalk Triple-tube swivel-type Minor core loss corebarrel with mud or msf Discontinuities opened flush, or retractor barrel Decomposed granite Treifus or Mazier retractor Minor core loss. barrel Effective stress loss Hard rock Double-tube swivel-type Discontinuities opened corebarrel

Soil type

32

Chapter 8

Laboratory testing INTRODUCTION

Laboratory testing is part of the physical survey. As an integral part of site investigation, the need for laboratory tests will often dictate the type and frequency of sample to be taken, and will therefore control the method of forming boreholes. Thus the type of sampling requires a precognition of the soil conditions on site; this has had the effect of leading some writers to recommend at least two stages of field work, with the bulk of laboratory testing being carried out after specific sampling in the second phase of investigation. For routine work such a programme is impractical and rarely used, because of the increases in cost and time that it causes. If two phases of site and laboratory work cannot be included then the investigation must be more carefully planned. With provision for changes during field work, with close engineering supervision and with a knowledge of soil conditions on site based on a first-class desk study, it should be possible to avoid the use of two field investigations. Soil mechanics, although a -branch of engineering, is often imprecise. Since many problems cannot be solved with accuracy, either as a result of imperfect analytical techniques or complex ground conditions, the use of refined sampling and testing techniques has been questioned. Terzaghi and Peck (1948) have commented ‘ ... On the overwhelming majority of jobs no more than an approximate forecast is needed, and if such a forecast cannot be made by simple means it cannot be made at all’. But is this attitude always justified? Certain classes of structure are so costly and the consequences of their failure so serious that, whatever the soil conditions, no effort should be spared in making as accurate a prediction of performance as possible. Where routine jobs are concerned, individual judgement based on low-cost sampling and testing may well suffice in the majority of cases, but such a method has a serious drawback; it does not allow extension of engineering knowledge based on observation and comparison with good quality data. Routine jobs are much more numerous than those for which the cost and time required for accurate and specialist testing can be justified, but can an engineer afford not to develop his experience and can he now afford the consequences of failure? Brunel and Stephenson could do so, for in their day experimental data were almost non-existent in the field of soil mechanics and it could be expected that the almost exclusive use of personal judgement would inevitably lead to some failures. We can no longer enjoy such luxury. When making predictions about the behaviour of soil, two factors are most important. First, it is normally necessary to judge which elements of soil behaviour will be critical to the satisfactory performance of the structure. Since there are many different ways in which soil behaviour can adversely affect the performance of a structure, it is necessary to appreciate all those facets which may cause problems and then analyse each, however briefly, to determine which are the most critical. Secondly, it is important to appreciate the limits which can be placed on any aspect of soil behaviour, for example, what settlement is tolerable, and is this the total or differential movement? For example, when considering the suitability of a site for spread footings for a multistorey structure it might be necessary to look at the following aspects of design: 1. 2. 3. 4.

overall slope stability after the end of construction; stability of temporary slopes during foundation construction; temporary support requirements; amount of seepage inflow into excavations;

Laboratory Testing 5. 6. 7. 8. 9. 10.

effects of seepage and loss of ground on adjacent structures; settlement of surrounding ground due to groundwater lowering; maximum allowable foundation bearing pressure; predicted settlements of footings; time for consolidation to occur; and proposed dimensions and layout to keep differential settlements small.

In any one case, it is probable that only a small proportion of these problems would require the acquisition of soil parameters for solution. The quality of the data required would depend on the allowable limits set for the structure. Thus spread footings in weathered rock would not normally experience significant settlements, but if a raft with very little tolerance of differential settlement were considered then even these conditions might give difficulties. Examples of the very small tolerance to differential settlements of sugar silo raft foundations, where doming of 5 mm over a 23 m foundation diameter was the limit to avoid structural distress, have been considered by Burland and Davidson (1976), Kee (1974) and Connor (1980). Two factors affect the quality of soil testing data required for a satisfactory prediction of soil behaviour. The tests carried out must be appropriate for the acquisition of the required data, or their results must be empirically linked to the required soil parameters with sufficient precision for the required calculation. In addition, sampling and testing must be carried out using techniques and accuracy which will yield parameters which are representative of the bulk of the soil in situ. Bearing in mind the small proportion of the on-site soil which will be sampled, (Broms,(1980) suggests 1 in 1000000 by volume), it will never be feasible to obtain representative parameters when soil conditions are variable, however good or expensive the sampling and testing techniques. Under these circumstances only simple laboratory tests should normally be considered, anti field tests may provide more useful data. THE PURPOSE OF SOIL TESTING

In general, soil is tested in order to assess its variability and in order to obtain parameters for particular geotechnical calculations. These two distinct reasons for testing lead to very different testing programmes. Routine tests carried out to allow the soil on a site to be divided into groups should ideally be scheduled for an initial phase of testing. Subsequent more expensive and complex tests are normally carried out on soil which is thought to be representative of each group; the samples to be tested cannot be so well selected before the results of classification tests are known. For reasons of time and economy, this ideal scheme cannot normally be used. More complex tests require a longer test period. When testing is started at about the same time as samples start to arrive from site, the engineer initially may have to rely completely on soil descriptions for a division of the in situ soil. Soil classification is carried out in order to define a small number of different groups of soil on any site. Each soil group may consist of a stratigraphically defined geological unit. More often it may ignore geological boundaries because the essence of the soil group should be that materials within it have (or are expected to have) similar geotechnical properties. Particle size, plasticity and organic content may be more important to the geotechnical engineer than time of deposition. The three main tools used to classify soil are soil description, particle size distribution analysis and plasticity testing. AVAILABLE TESTS

This chapter sets out to describe individual test techniques in detail: texts such as Lambe (1951), Bishop and Henkel (1962), Akroyd (1964), Vickers (1978), Head (1980), Head (1982), Head (1986), BS 1377:1990 and ASTM Part 19, should be referred to for the methods used in each test. Soil tests are loosely brought into two groups in this section; the first provides information to allow the

2

Site Investigation classification of soil into arbitrary groups while the second includes all tests which provide parameters which may be used in geotechnical calculation and design (Table 8.1). Table 8.1 Soil classification tests and test parameters Soil classification tests Tests for geotechnical parameters Sample description Strength tests (Discussed in Chapter 2) Particle size distribution Stiffness tests tests Plasticity tests Consolidation tests Compaction tests Seepage and permeability tests Specific gravity tests This division is not conventional. Normally plasticity tests, particle size distribution and specific gravity tests are know as soil classification tests (for example, see Head (1980) or BS 1377:part 2:1990). The British Standard used for soil testing for many years was BS 1377:1975. BS 1377:1975 comprised a single document which covered a wide range of tests for classification and geotechnical parameters. However, in certain areas the scope of the old British Standard was limited. For example, when it was written effective stress strength tests were not considered routine in most commercial laboratories and hence the description of such tests were omitted from the standard. BS 1377:1975 has now been completely revised and is superseded by BS 1377:1990. The new British Standard is divided into nine separate parts: Part 1 General requirements and sample preparation Part 2 Classification tests Part 3 Chemical and electro-chemical tests Part 4 Compaction-related tests Part 5 Compressibility, permeability and durability tests Part 6 Consolidation and permeability tests in hydraulic cells and with pore pressure measurement Part 7 Shear strength tests (total stress) Part 8 Shear strength tests (effective stress) Part 9 In situ tests.

Soil classification tests Soil classification, although introducing a further stage of data acquisition into site investigation, has an important role to play in reducing the costs and increasing the cost-effectiveness of laboratory testing. Together with detailed sample description, classification tests allow the soils on a site to be divided into a limited number of arbitrary groups, each of which is estimated to contain materials of similar geotechnical properties. Subsequent more expensive and time-consuming tests carried out to determine geotechnical parameters for design purposes may then be made on limited numbers of samples which are selected to be representative of the soil group in question. Particle size distribution tests BS 1377:1990 gives four methods for determining the particle size distribution of soils (part 2, clauses 9.2—9.5). The coarse fraction of the soil (>0.06mm approximately) is tested by passing it through a

3

Laboratory Testing series of sieves with diminishing apertures. The particle size distribution is obtained from records of the weight of soil particles retained on each sieve and is usually shown as a graph of ‘percentage passing by weight’ as a function of particle size (Fig. 8.1).

Fig. 8.1 Typical particle size distribution. Two methods of sieving are defined in BS 1377 (part 2, clauses 9.2, 9.3). Dry sieving is only suitable for sands and gravels which do not contain any clay: the British Standard discourages its use, and since the exact composition of a soil will not be known before testing, it is not often requested. Wet sieving requires a complex procedure to separate the fine clayey particles from the coarse fraction of the soil which is suitable for sieving, as summarized below. 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.

Select representative test specimen by quartering and riffling. Oven dry specimen at 105—110°C, and weigh. Place on 20mm sieve. Wirebrush each particle retained on the 20mm sieve to remove fines. Sieve particles coarser than 20 mm. Record weights retained on each sieve. Riffle particles finer than 20mm to reduce specimen mass to 2kg (approx.). Weigh. Spread soil in a tray and cover with water and sodium hexametaphosphate (2 g/l). Stir frequently for 1 h, to break down and separate clay particles. Place soil in small batches on a 2mm sieve resting on a 63 m sieve and wash gently to remove fines. When clean, place the material retained in an oven and dry at 105—110°C. Sieve through standard mesh sizes between 20mm and 6.3 mm using the dry sieving procedure. Note weights retained on each sieve. If more than 150 g passes the 6.3mm mesh, split the sample by riffling to give 100—150g. Sieve through standard mesh sizes between 5mm and 63 tm sieve.

It is important that this procedure is closely adhered to. Inadequate dispersal of the clay particles, poor washing, the overloading of sieves, and insufficient sieving time can all lead to inaccurate results. In particular, extra time and care may be required to ensure full dispersion of clay lumps within the test specimen.

4

Site Investigation The particle size distribution of the fine soil fraction, between about 0.1 mm and 1 µm may be determined by one of two British Standard sedimentation tests (BS 1377:part 2, clauses 9.4, 9.5). Soil is sedimented through water, and Stokes’ law, which relates the terminal velocity of a spherical particle falling through a liquid of known viscosity to its diameter and specific gravity, is used to deduce the particle size distribution. Sedimentation tests make a number of important assumptions. Since Stokes’ law is used, the following assumptions are implied (Allen 1975). 1. The drag force on each particle is due entirely to viscous forces within the fluid. The particles must be spherical, smooth and rigid, and there must be no slippage between them and the fluid. 2. Each particle must move as if it were a single particle in a fluid of infinite extent. 3. The terminal velocity must be reached very shortly after the test starts. 4. The settling velocity must be slow enough so that inertia effects are negligible. 5. The fluid must be homogeneous compared with the size of the particle. Since Stokes’ law applies only in the laminar flow region, for Reynolds numbers of less than 0.2, it cannot be applied to large particles. For quartz spheres (Gs = 2.65) falling in water the critical diameter is 60 µm. Some idea of the minimum particle size that can be measured by sedimentation in water can be obtained by considering the relative displacements per unit time of a small particle due to Brownian motion and gravity settlement. For particles finer than 1 µm Brownian motion exceeds gravitational motion, but in reality since Brownian motion is extremely weak when compared with even the slightest convection current the minimum particle size measurable is about 2tm. High concentrations of particles in the fluid create a number of problems. Because of the rigidity of the particles, increasing concentrations result in increases in apparent viscosity of the suspension. Additional problems, occur owing to particle—particle interaction: a cluster of particles may have a much greater terminal velocity than the individual particle settling velocities and at volume concentrations as low as 1% the suspension may settle en masse, apparently giving a coarser size distribution. High volume concentrations are also associated with the upflow of displaced fluid, causing an over-estimation of the fines content. Vickers (1978) expresses the view that provided that the concentration of soil is maintained at less than 50 g per 1000 ml and the container used for sedimentation is larger than 50mm dia., errors will generally be negligible. Allen (1975) indicates that concentrations should be less, than 1% by vol. (i.e. about 25g per l000ml). In addition to the assumptions and problems discussed in detail above, the nature of soil particles causes particular inaccuracies in sedimentation testing. First, the methods of preparation (i.e. mechanical agitation) may modify the particle size distribution. Secondly, the density of each soil particle will not equal the average specific gravity of the soil particles times the density of water: clay particles will contain adsorbed or absorbed water giving particle densities which may approach one half of the calculated value. Finally, few soil particles will be spherical. Clay particles will tend to be platy and will not drop vertically, and indeed may not be capable of achieving steady motion. Two techniques are available for sedimentation. The British Standard (BS 1377:1990) prefers the use of a fixed-depth pipette (BS 3406:part 2) (sometimes referred to in soil testing literature as an Andreasen pipette) to sample the soil — water mix at a depth of 10cm below the fluid surface at regular intervals after the test has been started by evenly distributing the soil in the water. The rate at which the suspension is drawn into the pipette is most important. In analysis, sampling is assumed to take place instantaneously, but the rapid withdrawal of a sample tends to give a finer distribution. BS 3406 recommends a 20s sampling time, while BS 1377 uses a l0s sampling time. The weight of soil left at that depth after a known time is determined by oven drying the sample, and Stokes’ law is then used to deduce the maximum particle size that can be left at that level.

5

Laboratory Testing

An alternative technique, which requires less sophisticated glassware, uses the hydrometer to determine the density of the soil — water mix at some depth. This method is less accurate in principle than the pipette, because the hydrometer does not measure density at a fixed point below the surface of the fluid, but determines an average value over the depth of its bulb. It is known that the pipette and hydrometer do not yield the same particle size distributions, and it is generally believed that the pipette is more accurate, but the effects of all of the inaccuracy of assumptions discussed above do not appear to have been assessed in absolute terms.

Plasticity tests The plasticity of soils is determined by using relatively simple remoulded strength tests. The plastic limit is the moisture content of the soil under test when remoulded and rolled between the tips of the fingers and a glass plate such that longitudinal and transverse cracks appear at a rolled diameter of 3 mm. At this point the soil has a stiff consistency. The liquid limit of a soil can be determined using the cone penetrometer or the Casagrande apparatus (BS 1377:1990:part 2, clauses 4.3, 4.5). One of the major changes introduced by the 1975 British Standard (BS 1377) was that the preferred method of liquid limit testing became the cone penetrometer. This preference is reinforced in the revised 1990 British Standard which refers to the cone penetrometer as the ‘definitive method’. The cone penetrometer is considered a more satisfactory method than the alternative because it is essentially a static test which relies on the shear strength of the soil, whereas the alternative Casagrande cup method introduces dynamic effects. In the penetrometer test, the liquid limit of the soil is the moisture content at which an 80 g, 300 cone sinks exactly 20 mm into a cup of remoulded soil in a 5s period. At this moisture content the soil will be very soft. Figure 8.2 shows the cone penetrometer and Casagrande cup. When determining the liquid limit with the Casagrande apparatus, the base of the cup is filled with soil and a groove is then made through the soil to the base of the cup. The apparatus is arranged to allow the metal cup to be raised repeatedly 10mm and dropped freely on to its rubber base at a constant rate of two drops per second. The liquid limit is the moisture content of a soil when 25 blows cause 13mm of closure of the groove at the base of the cup. The liquid limit is generally determined by mixing soils to consistencies just wet and dry of the liquid limit and determining the liquid limit moisture content by interpolation between four points (Fig. 8.3). BS 1377:part 2:1990, clause 4.6 provides factors which allow the liquid limit to be determined from one point (Clayton and Jukes 1978). The plastic limit test relies heavily on the skill of the operator, and is almost entirely subjective despite attempts by the British Standard to define procedure rigidly. The Casagrande cup method of determining the liquid limit is also rather operator dependent, and in addition suffers from apparatus maintenance problems. These two tests were subject to a comparative testing programme carried out in the UK and reported by Sherwood in 1970. The repeatability of these tests between over 40 laboratories in the UK was tested and gave the results listed in Table 8.2. The range of results reported for these tests is rather alarming, particularly in view of the fact that it was known by the participating organizations that their results would be compared with those of rival organizations. Sherwood (1970) commented that TRRL attempts to assess the amount of error attributable to defective or worn apparatus in the liquid limit test indicated that the majority of error was due to operator technique. This certainly agrees with our observations which include one of a 15% moisture content error in determining the liquid limit using the Casagrande apparatus as a result of incorrect frequency of drop. When considering the plastic limit test it is surprising that any agreement between laboratories exists. The amount of finger pressure used and the shape of the tips of fingers varies to a great extent and, in addition, operators frequently do not carry out the test using the tips of the fingers (as specified in the British Standard) since these are eminently unsuited to the task.

6

Site Investigation

Fig. 8.2 Casagrande cup and cone penetrometer for liquid limit testing. Table 8.2 Results of comparative testing programme Soil B Soil G Soil W Plastic limit (%) Mean 18 25 25 Range 13—24 18—36 20—39 S.D. 2.4 3.2 3.1 Coefficient of variation 13.1 12.8 12.7 Liquid limit (%) (Four-point method) Mean 34 69 67 Range 29—38 59—84 55—85 S.D. 2.4 5.2 5.3 Coefficient of variation 7.1 7.5 7.9

7

Laboratory Testing

Fig. 8.3 Liquid limit result by four-point cone method. The definitive method for the determination of liquid limit is the cone penetrometer. Operator technique can affect this test, particularly since it has been observed that long resting periods, after initially mixing the soil approximately to its liquid limit stage and before carrying out the test, tend to give higher results. (BS 1377:part 2 1990, clause 4.3 Note Three attempts to eliminate this effect by specifying a 24 h rest period between initial mixing of the soil with water, and carrying out the liquid limit test.) The requirement that each part of the test must be repeatable within fixed limits (if observed) however, leads to a much improved result. Tests reported by Sherwood and Ryley (1968), before the introduction of the test as a British Standard, indicated that ‘within laboratory’ variability is much reduced by the cone penetrometer method. The effects of operator technique between test houses are not known. Plasticity tests are widely used for classification of soils (Fig. 8.4) into groups on the basis of their position on the Casagrande chart (Casagrande 1948), but in addition they are used to determine the suitability of wet cohesive fill for use in earthworks, and to determine the thickness of sub-base required beneath highway pavements (Road Research Laboratory 1970). The results of wrong decisions in the latter two cases are likely to be much more serious than in the former case; test results from Sherwood (1970) indicate that single plasticity tests, or more than one plasticity test carried out by the same ‘biased’ operator cannot be used for these purposes.

Fig. 8.4 Casagrande plot showing classification of soil into groups.

8

Site Investigation

The extensive use of plasticity testing can be most rewarding, but the low levels of accuracy coupled with high cost tend to discourage use. At the present time liquid and plastic limit tests carried out to the British Standard in the preferred manner will normally take 48—72h to complete, allowing only for resting periods after mixing, and for oven-drying. The result of attempts to improve reproducibility has been a complexity of procedure which has increased expenses as Table 8.3 shows. Table 8.3 Relative costs of plasticity tests Cost of typical tests, divided by cost of one set of three 38mm dia. specimens tested in undrained Plasticity tests triaxial Plasticity tests compression (1995) Four-point cone penetrometer liquid limit test, plastic limit, plasticity index and 0.85 natural moisture content BS 1377:part 2, clauses 3, 4.3, 5 Four-point Casagrande liquid limit test, plastic limit plasticity index and natural 0.79 moisture content BS 1377:part 2, clauses 3, 4.5, 5 One-point Casagrande liquid limit test, plastic limit, plasticity index and natural 0.60 moisture content BS 1377:part 2, clauses 3, 4.6, 5 The low level of repeatability of the plastic limit test and the high cost and time-consuming nature of the four-point cone penetrometer liquid limit test make these tests unsuitable for construction control or for soil grouping. Clayton and Jukes (1978) have considered the possibility of a one-point cone penetrometer liquid limit, and concluded that such a test could provide a cheap but relatively accurate alternative to the one-point Casagrande method.

Compaction tests British Standard BS 1377: 1990:part 4 provides three specifications for laboratory compaction: 1. 2.5 kg rammer method; 2. 4.5 kg rammer method; and 3. vibrating hammer method for granular soils. Compaction has been defined as ‘the process whereby soil particles are constrained to pack more closely together through a reduction in the air voids, generally by mechanical means’ (Road Research Laboratory 1952). Compaction is therefore a rapid process which does not normally involve a significant change in moisture content. Laboratory compaction tests are intended to model the field process, and to indicate the most suitable moisture content for compaction (the ‘optimum moisture content’) at which the maximum dry density will be achieved for a particular soil. The 2.5 kg rammer method is derived from the work of Proctor (1933) which introduced a test intended to be relevant to the compaction techniques in use in earthfill dam construction in the USA in the 1930s. The test subsequently became adopted by the American Association of State Highway Officials (AASHO), and was known as the Proctor or AASHO compaction test. In the original test a mould of capacity 1/30 ft3 with an internal diameter of 4 in. was filled with soil at a fixed moisture content in three approximately equal layers. Each layer was compacted by 25 blows of a 2 in. dia. 5.5 lb rammer dropping through a height of 12 in. After compaction, the soil was trimmed to the level of the top of the mould, and the wet weight of soil and

9

Laboratory Testing its moisture content determined. The process was repeated for several increasing moisture contents, and a compaction curve (i.e. dry density as a function of moisture content) was obtained (Fig. 8.5).

Fig. 8.5 Compaction curves. Subsequent tests have been developed either to model advances in compaction plant, or as a result of metrication. The Modified AASHO test was developed during World War II to model the heavier standard of field compaction in use during airfield construction. A comparison of these tests is given in Table 8.4. The vibrating hammer compaction test was introduced in the 1967 revision of BS 1377. This test uses an electrical vibrating hammer with a tamper of approximately the same diameter as the mould (c.f. 145mm and 152mm); the electrical hammer is required to consume between 600 and 750W with an operational frequency between 25 and 45 Hz, and has a dead weight of between 300 and 400 N. Because of the limited size of the moulds in use, laboratory compaction tests require the exclusion of coarse soil particles. The conventional non-vibratory compaction tests that were covered by the 1975 British Standard made use of 1 litre moulds, necessitating the removal of particles held on the 20mm sieve. In the revised 1990 British Standard the specification for these tests have been extended to include soils with coarse gravel-size particles (BS 1377:part 4:1990, clauses 3.4, 3.6). These compaction tests are suitable for soils containing no more than 30% by mass of material retained on the 20mm sieve, which may include some particles retained on the 37.5 mm sieve. When compared with the conventional tests (clauses 3.3 and 3.5) where coarse gravel-size particles are removed it w1l be seen from the table above that a larger mould (CBR mould) together with a greater number of blows per layer (62) are specified. The vibrating hammer test (BS 1377:part 4:1990, clause 3.7) uses the CBR mould which is suitable for coarse gravel-size particles up to 37.5 mm in diameter. 10

Site Investigation

Test Proctor Modified AASHO BS 1377: part 4: 1990 Clause 3.3 Clause 3.4

Clause 3.5 Clause 3.6

Clause 3.7

Soil type

Particles up to mediumgravel size Soils with some coarse gravel-size particles Particles up to mediumgravel size Soils with some coarse gravel-size particles

Table 8.4 Comparison of compaction tests Height of Weight of No. of Blows/ drop rammer layers layer (in) (mm) (lb) (kg) 3 25 12 305 5.5 2.5 5 25 18 457 10 4.55

Volume of mould (ft3) (cc) 1/30 944 1/30 944

3

27

300

2.5

1000

3

62

300

2.5

CBR mould c. 2300

5

27

450

4.5

1000

5

62

450

4.5

CBR mould c. 2300

3

Vibratory 25-45 Hz

Vibratory

c. 30—40

CBR mould c. 2300

The repeatability of the 2.5 kg and 4.5 kg rammer methods of compaction, between laboratories, has been discussed by Sherwood (1970). Typically, the maximum variation of reported optimum moisture content was 4—8%, with reported maximum dry densities varying by up to 0.19 Mg/rn3. Inaccuracies of this magnitude make the tests unsuitable for either design or control, even if it is assumed that their results are relevant to field compaction conditions. In the UK, compaction tests are used for a variety of purposes. In their simplest application they are used to determine the optimum moisture content and maximum dry density expected of a soil; as a result, soils which have moisture contents widely different from the laboratory optimum may not be used in the construction of a fill and material which is found by in situ density measurement (see BS 5930:1981) to have a dry density considerably lower than the laboratory maximum (say less than 95%) may have to be removed from a fill and recompacted. The low level of repeatability and timeconsuming nature of compaction tests make them unsuitable for control tests, but in addition there is little evidence to suggest that their results give some optimum condition for the soil. Different types of soil react in very different ways to each type of roller. It is commonly known that increasing levels of compactive effort tend to produce higher maximum dry density values in conjunction with progressively lower optimum moisture contents, but results from Foster (1962) show that the ‘lines of optimums’ developed in field compaction trials with different plant are not coincident (Fig. 8.6). The objects of field compaction are to obtain sufficient strength, eliminate collapse, and reduce compressibility of fill to an acceptable level; it is doubtful if these aims can be achieved by the limited use of an empirical test with poor repeatability. This may explain the increasing use of specifications which either define the method of compaction in the field, or limit the air void content of the fill after compaction.

11

Laboratory Testing

Fig. 8.6 Lines of optimum moisture content/maximum dry density for laboratory compaction methods and two types of field compaction (Foster 1962). Particle density (specific gravity) determination Specific gravity values for a soil are not normally used strictly for classification purposes, but are used in the calculation and interpretation of other test results. The specific gravity tests specified in the British Standard (BS 1377:part 2:1990, clause 8) are relatively simple and are based upon determination of the dry weight of a sample of the soil, and the weight of the same sample plus water in a container of known volume. The volume of the container is obtained by weighing the container empty, and full of water. The main problems in conducting the test are of accurate weighing, and complete removal of the air from the soil after the addition of water. The method still used by most test houses to determine the particle density of fine-grained soil utilizes a 50m1 density bottle (BS 1377:part 2:1990 clause 8.3). Unfortunately there is no simple means of knowing when all the air has been removed from the bottle and hence the soil must be de-aired under vacuum. The use of de-aired water will help but it is still necessary to leave the sample in the density bottle under vacuum for several hours. The major difficulty with this test is the provision of a satisfactory vacuum and measuring the length of time required to remove the air completely. These factors can clearly lead to errors in specific gravity determinations. Krawczyk (1969) found that the difficulties of de-airing the soil could be overcome by shaking the sample instead of placing it under vacuum. The advantages of shaking are that the shaking action is easily standardized and the removal of air is more rapid than by the application of a vacuum. Krawczyk proposed that the test should be carried out in a 1 litre gas jar to make the same test suitable for fine-, medium- and coarse-grained soils and the shaking action provided by an end-over-end shaker. This alternative method has been included in the British Standard (BS 1377:part 2:1990, clause 8.2) and should be treated as the preferred method, since in providing a more reliable technique of de-airing the soil it yields more repeatable results.

12

Site Investigation

Results quoted by Sherwood (1970) for three clays tested with especial care to de-air by eight Road Research Laboratory operators are compared in Table 8.5 with the values from some 30 other test houses: they indicate that the specific gravity of British clays may be considerably higher than the 2.65—2.70 values typically expected by experienced engineers. Table 8.5 Comparison of particle density results RRL results Other test houses Clay type Mean Mm. Max. Mean Mm. Max. Bagshot beds 2.70 2.69 2.72 2.66 2.54 2.81 Gault clay 2.75 2.74 2.77 2.70 2.58 2.84 Weald clay 2.79 2.78 2.81 2.71 2.60 2.85 The results of particle density tests are used in the interpretation of sedimentation test results, to check the results of laboratory compaction tests (BS 1377:1975, clauses 4.1.4, 2.1), and to find the voids ratios of samples during consolidation tests. The test results quoted in Table 8.5 indicate a typical error in particle density determination of about 0.05. Incorrect particle density values affect the position of the voids ratio vs. logarithm of pressure plot for an oedometer consolidation test but they do not affect the values of the coefficients of consolidation (cv) or compressibility (mv). A change in particle density leads to a different particle size distribution from the sedimentation test, but the difference is not large and is probably considerably less than the effects of natural soil variability or the assumptions involved in the test. The major problem arising from an incorrect particle density determination is that of the credibility of compaction tests carried out on the same soil. A low particle density value will push the zero air voids line on a dry density/moisture content plot down and to the left, and may show compaction test results to be apparently impossible (and therefore inaccurate) as they cross over the zero air voids line. Tests for geotechnical parameters A wide range of tests has been used to determine the geotechnical parameters required in calculations for example, of bearing capacity, slope stability, earth pressure and settlement, but as testing techniques have changed some tests have been abandoned. Geotechnical calculations remain almost entirely semi-empirical in nature; it has been said that when calculating the stability of a slope one uses the ‘wrong’ slip circle with the ‘wrong’ shear strength to arrive at a satisfactory answer. For this reason testing requirements differ considerably from region to region. In Scandinavia the in situ vane is widely used to determine the undrained shear strength of clays while in Britain this parameter is normally determined using the unconsolidated undrained triaxial compression test. Bearing in mind the Norwegian Geotechnical Institute’s experience in applying Scandinavian techniques to the design of embankments in Asia, some caution should be exercised in introducing familiar techniques to unfamiliar ground conditions. Clearly each region develops its own testing techniques and comes to appreciate the necessary ‘factor of safety’ applicable to each type of calculation and each method of obtaining parameters. This section relates to laboratory practice in the UK at the time of writing.

Strength tests The principal tools available for strength determination in a good UK geotechnical testing laboratory are the California Bearing Ratio (CBR) apparatus, the Franklin Point Load Test apparatus (Franklin et

13

Laboratory Testing al. 1971; Broch and Franklin 1972), the laboratory vane apparatus and various forms of direct shear and triaxial apparatus. California bearing ratio (CBR) test The CBR and Franklin point load tests are empirical in nature. The CBR test is primarily used to assess the strength of materials used in or beneath flexible highway or airfield pavements. The test may be carried out in situ, or in the laboratory: BS 1377:part 4:1990, clause 7 gives a detailed description of the British Standard test method while the Road Research Laboratory publication (1952) describes the development of the test and some previous applications of its results. The CBR test was specifically developed by the California State Highway Department for the evaluation of sub-grade strengths in the investigation of existing pavements of known performance in use (Porter 1938, 1942). This led to an empirical method of pavement design. The test is carried out by forcing a standard plunger (approximately 50mm dia.) into the soil at a more or less constant rate of 1.25 mm/mm. Measurements of applied load and plunger penetration are made at regular intervals, and a curve is plotted for penetrations of up to 12.5 mm. Figure 8.7 shows the laboratory apparatus, and a typical result. The California Bearing Ratio is obtained by dividing the plunger loads at penetrations (after bedding correction) of 2.5 and 5.0 mm by the loads given at the same penetrations on a standard crushed stone. The loads given by the soil under test are expressed as percentages of the standard load, and the highest value is taken as the CBR value for design.

Fig. 8.7 CBR apparatus and typical result.

14

Site Investigation

The CBR test primarily involves shear deformation of the soil beneath the plunger, but its results cannot be accurately related to any of the fundamental shear strength parameters. Its use is therefore restricted to the design of road and airfield pavements. Because of the empirical nature of such designs it is of the utmost importance that the test is carried out precisely in the manner used to develop the particular design method in use. In the UK the CBR test is no longer widely used, because pavement design carried out based on the observed performance of pavements in the UK (Road Research Laboratory 1970) allows the CBR value to be obtained from particle size or plasticity index. Apart from a limited amount of testing to check the quality of sub- grade during construction, the only other use is to determine the strength of granular sub-base (Department of Transport 1976). Franklin point load test The Franklin point load test (Broch and Franklin 1972; ISRM 1985) was developed at Imperial College, London to provide a quick and reliable measurement of the strength of unprepared rock core samples, both in the field and the laboratory. The apparatus consists of a small loading frame which is activated by an hydraulic hand pump and ram (Fig. 8.8a). Rock core is placed between pointed platens of standard dimensions and loaded until failure occurs. The point load strength index:

Is =

P D2

(8.1)

where P = force required to break the specimen, and D = distance between the platen contact points. The results give a measure of the tensile strength of the rock. Tests may most reliably be carried out across the core diameter, but results can also be obtained when discs of core are loaded axially. Under this latter condition, corrections to the point load strength index will be required which will depend on the aspect ratio of the specimen (Broch and Franklin 1972). Under extreme circumstances, when only irregular lumps of rock are available, the test can be carried out along the shortest axis of the lump, but results will be less reliable. The value of diametral point load strength has been shown by Broch and Franklin to be dependent on the core size, with larger diameter cores giving smaller values of point load index. It has therefore been proposed that a standard classification be adopted by correcting all values to a reference diameter of 50mm. A correction chart for this purpose is given in Fig. 8.8b, based on the results of tests on five rock types at diameters between 10 and 80 mm.

Laboratory vane test The principles involved in the vane test are discussed in Chapter 8, under ‘In situ testing’. Whilst the field vane typically uses a blade with a height of about 150 mm, the laboratory vane is a small-scale device with a blade height of about 12.7mm and a width of about 12.7 mm. The small size of the laboratory vane makes the device unsuitable for testing samples with fissuring or fabric, and therefore it is not very frequently used. The laboratory vane test is described in BS 1377 :part 7:1990, clause 3.

Direct shear test The vane apparatus induces shear along a more or less predetermined shear surface. In this respect the direct shear test carried out in the shear box apparatus (Skempton and Bishop 1950) is similar. Figure 8.9 shows the basic components of the direct shear apparatus; soil is cut to fit tightly into a box which may be rectangular or circular in plan (Akroyd 1964; Vickers 1978; ASTM Part 19; Head 1982; BS 1377:1990), and is normally rectangular in elevation. The box is constructed to allow displacement along its horizontal mid-plane, and the upper surface of the soil is confined by a loading platen through which normal stress may be applied. Shear load is applied to the lower half of the box, the upper half being restrained by a proving ring or load cell which is used to record the shear load. The sample is not sealed in the shear box; it is free to drain from its top and bottom surfaces at all times.

15

Laboratory Testing The cross-sectional area over which the specimen is sheared is assumed to remain constant during the test.

Fig. 8.8 Point load apparatus (Broch and Franklin 1972; Brown and Phillips 1977). The direct shear test has been used to carry out undrained and drained shear tests, and to determine residual strength parameters. Morgenstern and Tchalenko (1967) reported the results of optical measurements on clays at various stages during the direct shear test, and it is clear that at peak shear stress and beyond, failure structures (Reidels and thrust structures) are not coincident with the supposed imposed horizontal plane of failure. In addition, the restraints of the ends of the box create an even more markedly non-uniform shear surface. Since the direction of the failure planes, the magnitude and directions of principal stresses and the pore pressure are not determinable in a normal shear box experiment, its results are open to various interpretations (Hill 1950), and this test is now rarely used to determine undrained or peak effective strength parameters. Triaxial tests may be performed more conveniently and with better control. In the UK, shear box tests are now used mainly to determine residual shear strength parameters for the analysis of pre-existing slope instability (Skempton 1964; Skempton and Petley 1967). In this application the technique of cut-plane shear box testing described by Petley (1966) gives results which have been found to be satisfactory, based on back-analysis (for example see Foster (1980)). A specimen of clay is placed in the shear box and allowed to swell for 24h under the weight of the load hanger. Following this, the specimen is consolidated under the required normal pressure and measurements of vertical compression are made. The two halves of the box are then separated

16

Site Investigation sufficiently to allow a cheese wire to cut smoothly through the specimen. The two halves of the specimen are then separated, and the soil surfaces smoothed by rubbing a glass plate lubricated by distilled water over the surfaces. When smooth, the lower half of the soil specimen is raised by packing it with a few layers of filter paper: the box is reassembled, and after applying a normal stress, the specimen is subjected to large displacements on the preformed shear surface by repeatedly reversing the travel of the box. The maximum shear stress obtained for each stage of shearing should be plotted against the logarithm of cumulative displacement, and shearing should continue until this curve levels out. The lowest maximum shear stress values (in the final shear stage of each test) are plotted against their imposed normal stresses to obtain the residual effective strength parameters (c'r (normally zero) and φ'r) for a soil. Typical values are given in Table 8.6.

Fig. 8.9 Bishop direct shear box. A better form of test to find residual parameters is carried out on an annular specimen in the ring shear apparatus, described by Bishop et al. (1971). Because of its cost and complexity this apparatus has failed to find a place in site investigation testing laboratories, but a simpler form of ring shear test described by Bromhead (1979) has been adopted by BS 1377:part 7:1990, clause 6. The simplified ring shear test is carried out on an annular specimen of remoulded clay 5mm thick, with internal and external diameters of 70mm and 100mm respectively. The specimen is confined radially between concentric rings and the vertical normal stress is applied via two porous bronze loading platens (Fig. 8.10). Relative rotary motion takes place between the confining rings (which are fixed to the lower loading platen) and the upper platen. This causes the sample to shear,, the shear surfaces forming close to the upper platen. The loading platens are roughened in order to prevent slip at the platen—soil interface. The upper platen reacts against two matched proving rings (or load cells) which provide a measurement of the torque transmitted through the soil specimen.

17

Laboratory Testing

Soil type

Table 8.6 Effective strength parameters for some UK soils Peak shear strength Residual shear strength wl (%) wp (%) c'r φ'r c'r φ'r (deg) (kN/m2) (deg) (kN/m2)

Sand and gravel Loose Medium dense Sand Loose Dense Silt Loose Dense Chalk Senonian, remoulded Cenomanian, intact Granular glacial till Oxford clay Weald clay Gault clay Unweathered London clay Weathered London clay

28

22

42

17

15—24 57 60—65 55 71 70—90

12—14 27 25—32 23 29 25—30

0 0

36—42 40—48

0 0

30—34 37—43

0 0

28—32 30—34

0

30—34

600— 1000 0—40 172 8 53 125 16—20

32—37 35—42 28 22 22 26 19—21

4 0 0 0 0

13 9—15 18 10 9—14

The main advantage of the Bromhead ring shear apparatus is that it is relatively simple but still allows an infinite relative displacement without the necessity of reversing the direction of relative motion along the shear plane developed in the soil sample. Preparation of a sample for the test involves kneading remoulded soil into the annular cavity formed by the confining rings and the lower platen. Surplus material is struck off such that the top surface of the soil is level with the top of the confining rings. With the upper platen placed on top of the soil, and the surrounding water bath filled to prevent evaporation during the test, the specimen is consolidated under a normal stress by weights applied to the load hanger.

Fig. 8.10 Bromhead ring shear apparatus (Bromhead 1979).

18

Site Investigation

The rate of shearing during drained direct shear testing must be slow enough to ensure that no excess pore water pressure exists on the failure plane by the time shear strength measurements are to be made. In practice it is normal to shear the specimen slowly enough so that excess pore pressures are insignificant when the peak shear strength is developed. The time to failure can be determined from the time/consolidation data obtained before the start of the shear stage, from which the coefficient of consolidation of the soil can be obtained, because: cv =

Tv d 2 t

(8.2)

where Tv = time factor at some specified percentage of consolidation, d = average drainage path distance for the specimen, and t = time taken for the specimen to reach the specified percentage of consolidation (see, for example, Terzaghi (1923, 1943)). Taylor and Merchant’s method (Taylor and Merchant 1940;BS 1377:part 5:1990) can be used to determine the point on a plot of compression as a function of square root of time at which 90% consolidation has occurred. At 90% consolidation of time factor, Tv = 0.848. Gibson and Henkel (1954) have expressed the minimum time to failure in the shear box as: tf =

h2 2c v (1 − U c )

(8.3)

where h = half the specimen height, cv = coefficient of consolidation, determined above, and U c = mid-plane pore-pressure dissipation ratio. A minimum dissipation ratio of 0.95 (i.e. 95% dissipation of excess pore pressures) is normally used, and the rate of deformation in the experiment is then calculated using an estimated value of deformation at failure, based on experience. If, subsequently, failure occurs before the required minimum time to failure then the results of the test may be invalid. BS 1377:1990 recommends that the time to failure should be based on the elapsed time for 100% consolidation (t100) using the relationship:

tf = 12.7t100

(8.4)

A similar approach based on t1 is also recommended by the 1990 British Standard for determining the rate of displacement for ring shear tests. However, most ring shear tests are carried out at a speed of 0.048% mm when using the Bromhead apparatus. The British Standard suggests in a note in part 7, section 6.4.5.1 that this speed has been found satisfactory for a large range of soils. Faster rates are likely to disrupt the sliding surface and result in erroneous values of φr. The residual angle of shearing resistance φr varies with the effective normal stress (σ'n) acting on the sliding surface (Bishop et al. 1971; Bromhead 1979; Lupini et al. 1981). For London clay, Bishop et al. found that φr was 14° at low effective normal stress (σ'n< 30kPa) and reduces to less than 9° at effective normal stresses greater than 100 kPa. Such variations in residual strength make it necessary to measure φr at a range of effective normal stresses. The procedure adopted for use with the Bromhead ring shear apparatus by most commercial laboratories is as follows. 1. Place the initial (lowest) load on the hanger and, with the gear disengaged, create a shear surface in the upper part of the specimen by rotating the handwheel through one or two revolutions. After rotating the handwheel ensure that any load on the proving rings or load cells is taken off by reversing the direction of handwheel rotation.

19

Laboratory Testing 2. Take a set of initial readings, engage the gear, start the motor and shear for several hours until a constant torque measurement is achieved. Generally, shearing overnight has been found satisfactory. 3. Place the next load on the hanger and continue to shear whilst taking readings at regular intervals until the torque remains constant for 20 mm. The whole stage should take about 1 h. 4. Repeat stage (3) at least three times with increasing normal loads. 5. When the maximum load is reached, stop the motor and manually reverse the rotation until a zero torque reading is achieved. If the torque is not taken off, the release of proving ring energy into the specimen causes disruption of the shear surface. 6. Reduce the load on the hanger back to the initial (lowest) load and commence shearing again until a constant torque reading is achieved. 7. Compare the torque reading from stage (6) with that at the end of stage (2). If the readings are the same, the test is complete; if they differ significantly then the test must be repeated. This procedure, although in common usage, is not given in the 1990 British Standard. A typical ring shear test performed in the above manner should take about 24 h to complete. Typical values of normal stress used in such tests include 25, 50, 75 and 100 kPa.

Triaxial test The triaxial apparatus has been described in great detail by Bishop and Henkel (1962). The test specimen is normally a cylinder with an aspect ratio of two, which is sealed on its sides by a rubber membrane attached by rubber ‘O’ rings to a base pedestal and top cap (Fig. 8.11). Water pressure inside the cell provides the horizontal principal total stresses, while the vertical pressure at the top cap is produced by the cell fluid pressure and the ram force. The use of an aspect ratio of two ensures that the effects of the radial shear stresses between soil, and top cap and base-pedestal are insignificant at the centre of the specimen.

Fig. 8.11 Triaxial cell.

20

Site Investigation The triaxial apparatus requires one or two self-compensating constant pressure systems, a volume change measuring device and several water pressure sensing devices. The ram force may be measured outside the cell using a proving ring, but most modern systems now use an internal electrical load cell mounted on the bottom of the ram. The ram is driven into the triaxial cell by an electrical loading frame which will typically have a capacity of 5000 or 10000 kgf and is capable of running at a wide range of constant speeds; triaxial tests are normally carried out at a controlled rate of strain increase. When this apparatus is used to measure strength the specimen is normally failed in triaxial compression, that is with the intermediate principal stress held constant and equal to the minor principal stress and with the major principal stress increased to bring about failure. Under these conditions the height of the specimen decreases during shearing. The three most common forms of test are: 1. the unconsoldiated undrained triaxial compression test, without pore water pressure measurement (BS 1377:part 7:1990. clause 8); 2. the consolidated undrained triaxial compression test, with pore water pressure measurement (BS 1377:part 8:1990, clause 7); and 3. the consolidated drained triaxial compression test, with volume change measurement (BS 1377:part 8:1990, clause 8). The unconsolidated undrained triaxial compression test is carried out on ‘undisturbed’ samples of clay in order to determine the undrained shear strength of the deposit in situ. Pore pressures are not measured during this test and therefore the results can only be interpreted in terms of total stress. Three test specimens, which may be either 38mm or 102mm dia. and will normally have an aspect ratio of 2, are extruded from a core and sealed using a rubber membrane, ‘O’ rings and top and bottom caps. Once a specimen is inside the triaxial cell, the cell pressure is increased to a predetermined value and the specimen is brought to failure by increasing the vertical stress; during this period regular readings of the ram load and specimen height decrease are made. The cell pressures used will normally increase by a factor of two between each of the three specimens, with the middle pressure approximately corresponding to the vertical total stress at the level of sampling in the ground. Thus for a sample taken from 5 m depth cell pressures of 50, 100 and 200 kN/m2 would be used. The rate of strain used during the test will normally be 2%/mm. This rate is based on the specifications given in the 1975 British Standard for the maximum strain (20%) and the maximum test duration (10mm.). However, BS 1377:1990 (part 7, clause 8) recommends that the rate of axial deformation should produce failure within a period of 5—15 mm. The recommendation concerning the maximum axial strain remains unchanged. If the same criteria for selecting a rate of strain are adopted using these recommendation the rate of strain should be 1.5%/mm. It should be pointed out that the undrained strength is not a fundamental property of the soil and the measured strength is sensitive to the rate at which the soil is sheared. It is therefore advisable to adopt the same rate for all tests of this type. Since the major total principal stress (acting in a vertical direction) is composed of two components, i.e.

σ1 = σ 3 +

P A

(8.5)

where σ3 = horizontal total stress (the cell pressure), P= ram force, and A = specimen cross-sectional area. The principal stress difference (or deviator stress), σ1- σ3, is simply equal to the ram force divided by the cross-sectional area. Because the test is carried out undrained, with no volume change allowed, the specimen diameter increases during the test. In order to calculate the cross-sectional area at any time during testing it is assumed that the specimen deforms as a right cylinder and so:

V= A0H0 = AH = AH0 (1 — εa)

(8.6) 21

Laboratory Testing

where V= specimen volume (constant), A0 and H0 are the original specimen area and height, A and H are the specimen area and height at some time during the test, and εa is axial strain at some time during the test. Thus A=A0/(l- εa). The results of the test are plotted as curves of principal stress difference against strain. For conditions of maximum principal stress difference (taken as failure) Mohr circles are plotted in terms of total stress. The average undrained shear strength should be quoted, and the failure envelope drawn tangential to the Mohr circles in order to find the undrained ‘cohesion intercept’ and undrained ‘angle of shearing resistance’. A correction should be applied to the measured maximum deviator stress to allow for the restraining effect of the membrane (BS 1377:1990). For a barrelling type of failure which occurs in a plastic soil the correction (σmb) is given by:

σ mb =

4Mε a (1 − ε a ) (kN/m 2 ) D

(8.7)

where M = compression modulus of the membrane material per unit width, εa = axial strain at failure, and D = initial diameter of specimen. The compression modulus of the membrane material, M, is assumed to be equal to its extension ‘modulus. The method by which the extension modulus is measured is described by Bishop and Henkel (1962) and Head (1982). In soils which exhibit brittle failure a different membrane correction may be necessary, although not mentioned in the British Standard. This correction is described by Head (1982). For soils of high strength, such as stiff clays, the effect of the membrane restraint is small and is often neglected. For soft and very soft clays the membrane effect can be significant and omission of the correction could lead to errors on the unsafe side. The membrane correction described above is deducted from the maximum measured deviator stress. The size of specimen tested in the undrained triaxial test can have a significant effect on the resulting shear strength (Bishop et al. 1965; Agarwal 1968; Marsland and Randolph 1977). While larger specimens may give parameters which are more relevant, for example, to slope stability calculation (for example, Skempton and La Rochelle (1965)) because of their inclusion of fissures or fabric, it is important to recognize that some empirical or semi-empirical design methods were specifically designed on the basis of undrained shear strengths measured on small diameter specimens. Strutted excavations (Peck 1969) and the adhesion on bored piles (Skempton 1959) are examples of this type of problem. The decision to test large diameter specimens can cause particular problems when, as is often the case, deposits are more variable in the vertical than the horizontal direction. Three 204mm high specimens cannot be taken from a standard 450mm long open-drive tube sample. To overcome the problems of shortage of material the aspect ratio of the specimens may be reduced to one by using lubricated end platens (Rowe and Barden 1964) or each specimen may be sheared at three cell pressure levels (Taylor 1950; Parry 1963; Anderson 1974). This latter technique is known as ‘multi-staging’, and has been found to be particularly useful in boulder clay materials where stone content makes the preparation of undisturbed specimens difficult, and test results from individual specimens typically give a large strength variation. Multistage tests are described in Head (1982) and BS 1377:part 7:1990, clause 9.

22

Site Investigation Peak effective strength parameters (c' and φ') may be determined either from the results of consolidated undrained triaxial compression tests with pore pressure measurement or from consolidated drained triaxial compression tests. The former test is normally preferred because it can be performed more quickly and therefore more economically. The consolidated undrained triaxial compression test is normally performed in several stages, involving the successive saturation, consolidation and shearing of each of three specimens. Saturation is carried out in order to ensure that the pore fluid in the specimen does not contain free air. If this occurs, the pore air pressure and pore water pressure will differ owing to surface tension effects: the average pore pressure cannot be found as it will not be known whether the measured pore pressure is due to the pore air or pore water, and at what level between the two the average pressure lies. Perhaps more importantly, the presence of air in the pore pressure measuring system can lead to time lags, which for relatively incompressible over-consolidated clay soils can be very significant. Bishop and Henkel (1962) quote theoretical times for 98% equalization of pore pressure for undisturbed 38 mm London clay specimens which vary from about 1 mm to 6h, depending on the compressibility of the pore pressure measuring system. Saturation is normally carried out by leaving the specimens to swell against an elevated back pressure. The use of a back pressure on dense specimens which are expected to dilate has the additional advantages of extending the range of applied stress for which pore pressure measurements can be made and, in drained tests, of preventing the formation of air locks in the triaxial pedestal and pipework leading to the specimen. Back pressure (which is simply an imposed pore pressure) is applied through a volume change gauge to the top of the specimen, while a cell pressure of slightly higher value is also applied. Both cell pressure and back pressure are normally increased in increments of about 50 kN/m2, allowing time for equalization at each stage. The degree of saturation can be expressed in terms of Skempton’s pore pressure parameter (Skempton 1954):

B=

∆u ∆σ 3

(8.8)

where ∆u = change in pore pressure for an applied cell pressure change of ∆σ3. For a saturated soil, B equals unity. In practice it has been found that B approximates to unity (say B ≥ 0.98) when a back pressure of 200—300 kN/m2 has been used on natural clays, but compacted samples may require back pressures of 400—800 kN/m2. Once a reasonable back pressure has been achieved the B value can be checked by measuring the response of the pore pressure to an applied cell pressure change. BS 1377:part 8:1990, clause 5 recommends that a value of B greater than or equal to 0.95 must be achieved before the specimen may be considered as fully saturated and the consolidation stage started. The consolidation stage of an effective stress triaxial test is carried out for two reasons. First, three specimens are tested and consolidated at three different effective pressures, in order to give specimens of different strengths which will produce widely spaced effective stress Mohr circles. Secondly, the results of consolidation are used to determine the minimum time to failure in the shear stage. The effective consolidation pressures (i.e. cell pressure minus back pressure) will normally be increased by a factor of two between each specimen, with the middle pressure approximating to the vertical effective stress in the ground. When the consolidation cell pressure and back pressure are applied to the specimen, readings of volume change are made using a volume change device in the back pressure line. The speed at which volume change takes place depends on the effective pressure increment, the coefficient of consolidation of the soil and the drainage conditions at the specimen boundaries. Normally, pore

23

Laboratory Testing pressure will be measured at the specimen base, with drainage to the back pressure line taking place through a porous stone covering the top of the specimen. The speed at which heavy clays consolidate and may be sheared can be significantly increased by the use of filter paper drains on the radial boundary of the specimen (Bishop and Henkel 1962). The coefficient of consolidation of the clay can be determined by plotting volume change as a function of the square root of time. Theoretical considerations indicate that the first 50% of volume loss during consolidation should show as a straight line on this plot. This straight line is extended down to cut the horizontal line representing 100% consolidation, and the time intercept at this point (termed ‘t100’ by Bishop and Henkel) can be used to obtain the coefficient of consolidation as shown below (in fact, t1 is equal to 4 x t50, and cannot equal the infinite time theoretically required for complete consolidation, see Fig. 8.12).

Fig. 8.12 Consolidation and shear stage results for a consolidated undrained triaxial compression test with pore pressure measurement. (Arrows denote principal stress ratio failure, (σ1'/σ3')max.) For drainage from one end only:

24

Site Investigation

cv =

πh 2

(8.9)

t100

For drainage from both ends and radial:

cv =

πh 2

(8.10)

100t100

where cv is the coefficient of consolidation of the clay and h is one-half of the specimen height. The filter drains used on the radial boundary of triaxial specimens do not cover their entire periphery, and are not infinitely permeable. Work by Bishop and Gibson (1964) indicates that the equations above may significantly under-estimate the coefficient of consolidation of more permeable clays, or silts, tested at high effective pressures, and it will therefore be unwise to use these results indiscriminately for other types of engineering calculation. A minimum time to failure during the shear stage is necessary, not only to allow for the time lag in the pore pressure measuring system, but also to allow equalization of pore pressures within the specimen. Friction between the specimen and the porous stones creates non-uniformity of stress and strain conditions between the centre and ends of the specimen, and as a result the pore pressures set up during an undrained test are non-uniform. Testing must be carried out slowly enough so that almost complete equalization of pore pressures at the centre and ends of the specimen takes place. For drainage from one end only: tf =

1.67 h 2 cv

(8.11)

0.071h 2 cv

(8.12)

For drainage from both ends and radial boundary: tf =

for 95% equalization of pore pressure in specimens of diameter, h, and height, 2h. The minimum time to failure (tf), or to the first valid effective stress readings if a stress path is required, can be obtained by combining the equations above to give:

tf = 0.53 tl00 tf = 2.26 tl00

for drainage from one end for drainage from the entire boundary

(8.13) (8.14)

The equations are strictly only valid if the major assumptions in their derivation are correct. It is assumed, inter alia, that the pore pressure differences are parabolic over the specimen height and are proportional to the applied load. When brittle failure is expected to take place over a narrow failure zone the rate of testing should be of the order of 10 times slower (La Rochelle 1960). Once consolidation is complete, the specimen may be isolated from the back pressure and the rate of vertical movement of the compression machine platen set. For this, the minimum time to failure is divided into the estimated axial sample deformation at failure (or at the time of the first valid readings). Soil will normally fail at axial strains of between 2 and 20%, and the actual figure used is largely based on experience of testing similar soil types. During the shear stage the vertical stress is

25

Laboratory Testing increased by the loading ram, and measurements are made at regular intervals of deformation, ram load and pore pressure. These are converted to graphs of principal stress difference (σ1- σ3) and pore pressure as a function of strain (Fig. 8.12), and failure is normally taken as the point of maximum principal stress difference. The effective stress Mohr circles are plotted for the failure conditions of the three specimens, and the gradient and intercept of a straight line drawn tangential to these circles defines the effective strength parameters c' and φ' (Fig. 8.13).

Fig. 8.13 Stress paths and Mohr circles at failure for a consolidated undrained triaxial compression test with pore pressure measurement. Effective stress triaxial tests are far less affected by sample size effects than undrained triaxial tests, but the problems of sampling in stoney soils still make multistage testing an attractive proposition under certain circumstances. The effectiveness of this technique in consolidated undrained triaxial testing has been reported by Kenney and Watson (1961), Parry (1968) and Parry and Nadarajah (1973). The consolidated drained triaxial compression test, with volume change measurement during shear is carried out in a similar sequence to the consolidated undrained test, but during shear the back pressure remains connected to the specimen which is loaded sufficiently slowly to avoid the development of excess pore pressures. The coefficient of consolidation of the soil is derived in the manner described above from the volume change measurements made during the consolidation stage. Gibson and Henkel (1954) found that the average degree of consolidation at failure is related to the time from the start of the test by the equation:

26

Site Investigation

tf =

h2 ηc v (1 − U f )

(8.15)

where 2h = specimen height, cv = coefficient of consolidation, and U f = average degree of consolidation at time t. For a specimen with an aspect ratio of 2, η equals 0.75 for drainage from one end only and 40.4 for drainage from both ends and the radial boundary. Thus, to achieve 95% consolidation: tf = 8.48 tl00 for drainage from one end only tf = 15.78 tl00 for drainage from both ends and the radial boundary

(8.16) (8.17)

Thus the shear stage of a drained triaxial test can be expected to take between 7 and 15 times longer than that of an undrained test with pore pressure measurement. 100mm dia. specimens of clay may require to be sheared for as much as one month. Once shearing is complete, the results are presented as graphs of principal stress difference and volume change as a function of strain, and the failure Mohr circles are plotted to give the drained failure envelope defined by the parameters cd' and φd' . The effective strength parameters defined by drained triaxial testing should not be expected to be precisely the same as those for an undrained test, since volume changes occurring at failure involve work being done by or against the cell pressure (Skempton and Bishop 1954). In practice the resulting angles of friction for cohesive soils are normally within 1—2°, and the cohesion intercepts are within 5 kN/m2. The results of tests on sands can vary very greatly (for example, Skinner 1969). Stiffness tests

From the 1950s through to the early 1980s there has been a preoccupation in commercial soil testing with the measurement of strength with less emphasis being paid to the measurement of detailed stress—strain properties such as stiffness. This is reflected in both the 1975 and the 1990 editions of BS 1377, both of which fail to consider the measurement of stiffness. During the last decade, two important parallel developments have taken place which have resulted in the measurement of stiffness being considered more important than that of strength in geotechnical design, particularly for sensitive structures. These developments are: 1. methods of measuring strain locally on laboratory test specimens have shown that the stress— strain behaviour of many soils and weak rocks is significantly non-linear with very high stiffness at the small strains operational around most engineering structures (Jardine et al. 1984); and 2. certain features of field measurements of ground deformation around full-scale structures, which could not be modelled using linear elastic theory, are resolved if non-linear formulations are used incorporating very high initial stiffness (Simpson et al. 1979). These developments have resulted in the application of finite element models in geotechnical design becoming commonplace, with stiffness parameters derived both from special laboratory stiffness tests and from field geophysics. In most soils any discontinuities such as fissures will generally have a stiffness that is similar to that of the intact soil such that the intact soil stiffness may be used to predict with reasonable accuracy ground deformations and stress distributions. This means that laboratory triaxial tests on good quality ‘undisturbed’ specimens may yield adequate stiffness parameters for design purposes. However,

27

Laboratory Testing conventional measurements of axial deformation of triaxial specimens, made outside the triaxial cell, introduce significant errors in the computation of strains. The conventional method of measuring axial deformation is shown schematically in Fig. 8.14a. The errors in the computation of strain that arise from this method of measurement result from the fact that apparatus is compliant; the load cell, porous stones, lubricated end platens (when used) and filter papers will all compress under increasing axial load (Baldi et al. 1988). Further errors are associated with bedding caused by lack of fit or surface irregularities at the interfaces between the specimen and loading surfaces (Daramola 1978; Burland and Symes 1982). Although the errors due to apparatus compliance can be evaluated with reasonable certainty by careful calibration, the bedding error can be very difficult to assess since its magnitude depends on the way in which the ends of the specimen are prepared. Thus the only way to obtain accurate determinations of axial strain is to carry out the measurement remotely from the ends of the specimen, and preferably on its middle third (Fig. 8.14b). This type of measurement is referred to as ‘local strain’ measurement. A comparison of local and axial strain measurements made on the same test specimen are shown in Fig. 8.15. It will be seen that the errors are greatest during the early stages of the test.

Fig. 8.14 Schematic diagram illustrating external and local strain measurement in the triaxial apparatus.

28

Site Investigation

Fig. 8.15 Comparison of local and external strains (Clayton and Khatrush 1986).

During the 1980s the accurate measurement of soil stiffness at small strains (
View more...

Comments

Copyright ©2017 KUPDF Inc.
SUPPORT KUPDF