31.12.2013 Views

an engineering geological characterisation of tropical clays - GBV

an engineering geological characterisation of tropical clays - GBV

an engineering geological characterisation of tropical clays - GBV

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

AN ENGINEERING GEOLOGICAL CHARACTERISATION OF TROPICAL<br />

CLAYS<br />

CASE STUDY: CLAY SOILS OF NAIROBI, KENYA<br />

Dissertation<br />

zur Erl<strong>an</strong>gung des Grades eines<br />

Doktors der Naturwissenschaften<br />

Vorgelegt von<br />

Simon W<strong>an</strong>yonyi Ot<strong>an</strong>do, M.Sc.<br />

aus Bungoma, Kenya<br />

Genehmigt von der<br />

Mathematisch-Naturwissenschaftlichen Fakultät<br />

der Technischen Universität Clausthal<br />

Tag der mündlichen Prüfung<br />

01.10.2004


Referent: Pr<strong>of</strong>. G. Reik, Ph.D.<br />

Korreferent: Pr<strong>of</strong>. Dr. H.-J. Gursky<br />

Dek<strong>an</strong>: Pr<strong>of</strong>. Dr. D. Mayer<br />

Die Arbeit wurde <strong>an</strong> der Abteilung für Ingenieurgeologie des Institutes für Geologie und<br />

Paläontologie der TU Clausthal <strong>an</strong>gefertigt.


i<br />

Contents<br />

Page<br />

Acknowledgements<br />

Summary (Abstract)<br />

Chapter 1. Introduction 1<br />

1.1 Scope <strong>of</strong> study 1<br />

1.2 Location <strong>of</strong> the study area 2<br />

1.3 Definition <strong>of</strong> study problem 2<br />

1.4 Summary <strong>of</strong> case problems 4<br />

1.5 Physiography 8<br />

1.6 Climate 11<br />

1.7 Method <strong>of</strong> survey 13<br />

Chapter 2. Previous works 15<br />

2.1 Summary 15<br />

Chapter 3. Geology 19<br />

3.1 Introduction 19<br />

3.2 Precambri<strong>an</strong> rocks <strong>of</strong> the Mozambique Belt 20<br />

3.3 Volc<strong>an</strong>ic rocks 21<br />

3.3.1 Kapiti phonolite 24<br />

3.3.2 Nairobi phonolite 25<br />

3.3.3 Nairobi trachytes 27<br />

Chapter 4. Field methods 30<br />

4.1 Introduction 30<br />

4.2 Investigation <strong>an</strong>d sampling scheme 30<br />

4.3 Cone penetration depth sounding 34<br />

4.4 Field v<strong>an</strong>e test 38<br />

4.4.1 Introduction 38<br />

4.4.2 Testing procedure 38<br />

4.4.3 Results <strong>of</strong> study 40<br />

4.4.4 Applications 43<br />

Chapter 5. Types <strong>of</strong> soil 45<br />

5.1 Superficial deposits 45<br />

5.2 Soils 46<br />

5.2.1 Red friable <strong>clays</strong> 48<br />

5.2.2 Black <strong>clays</strong> (black cotton soils) 53


ii<br />

Chapter 6. Chemical <strong>an</strong>d mineralogical <strong>an</strong>alysis 58<br />

6.1 Introduction 58<br />

6.2 X-ray fluorescence (XRF) studies 58<br />

6.2.1 Scope <strong>an</strong>d method 58<br />

6.2.2 Results 58<br />

6.3 X-ray diffraction (XRD) studies 59<br />

6.3.1 Scope <strong>an</strong>d method 59<br />

6.3.2 Results 60<br />

6.4 Sc<strong>an</strong>ning electron microscope (SEM) <strong>an</strong>alysis 67<br />

6.4.1 Scope <strong>an</strong>d method 67<br />

6.4.2 Results 68<br />

6.4.2.1 Nairobi phonolite 68<br />

6.4.2.2 Nairobi trachytes 70<br />

6.4.2.3 Black <strong>clays</strong> 71<br />

6.4.2.4 Red soils 75<br />

6.5 Import<strong>an</strong>ce <strong>an</strong>d influence <strong>of</strong> clay minerals on <strong>engineering</strong><br />

properties <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 77<br />

6.6 Carbon <strong>an</strong>d sulphur content <strong>of</strong> soils 78<br />

6.6.1 Method <strong>of</strong> <strong>an</strong>alysis 78<br />

6.6.2 Results 79<br />

6.6.3 Analysis <strong>of</strong> results 80<br />

Chapter 7. Laboratory soils index <strong>an</strong>d <strong>engineering</strong> properties determination 83<br />

7.1 Moisture content <strong>an</strong>d index tests 83<br />

7.1.1 Introduction 83<br />

7.1.2 Natural moisture content 83<br />

7.1.3 Atterberg limits <strong>an</strong>d consistency <strong>of</strong> clay soils 84<br />

7.1.4 Free swell test (after Gibbs & Holtz, 1956) 91<br />

7.1.5 Linear shrinkage 95<br />

7.2 Grain/ particle size <strong>an</strong>alysis 96<br />

7.2.1 Scope 96<br />

7.2.2 Procedure <strong>an</strong>d results 97<br />

7.3 Direct shear tests 104<br />

7.3.1 Scope 104<br />

7.3.2 Procedure 104<br />

7.3.3 Results 106<br />

7.3.4 Evaluation <strong>an</strong>d <strong>an</strong>alysis <strong>of</strong> results 110<br />

7.4 Oedometer consolidation tests 115<br />

7.4.1 Scope 115<br />

7.4.2 Theory <strong>of</strong> consolidation 115<br />

7.4.3 Consolidation coefficients <strong>an</strong>d other parameters 119<br />

7.4.4 Procedure 124<br />

7.4.5 Analysis <strong>of</strong> consolidation test data 125<br />

7.4.6 Results <strong>of</strong> consolidation tests 126<br />

7.4.7 Swelling pressure <strong>an</strong>d percentage swelling 138<br />

7.4.8 Evaluation <strong>an</strong>d application <strong>of</strong> results <strong>of</strong> oedometer consolidation,<br />

swelling pressure <strong>an</strong>d percentage swelling 152


iii<br />

Chapter 8. Distribution <strong>of</strong> index <strong>an</strong>d strength properties in black <strong>clays</strong> 155<br />

8.1 Natural moisture content 155<br />

8.2 Liquid limit 158<br />

8.3 Plasticity index 161<br />

8.4 Linear shrinkage 164<br />

8.5 Free swell 166<br />

8.6 Clay <strong>an</strong>d fine fraction 168<br />

8.7 Shear strength 173<br />

Chapter 9. Implications <strong>of</strong> index/ <strong>engineering</strong> properties <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils<br />

on construction practice 176<br />

9.1 Atterberg limits <strong>an</strong>d other index parameters 176<br />

9.2 Grain size distribution 177<br />

9.3 Shear strength parameters 177<br />

9.4 Consolidation-settlement characteristics 178<br />

9.5 Percentage swelling <strong>an</strong>d swelling pressure effects 180<br />

Chapter 10. Correlation <strong>of</strong> index properties 181<br />

10.1 Linear shrinkage <strong>an</strong>d plasticity index 181<br />

10.2 Liquid limit <strong>an</strong>d plasticity index 187<br />

10.3 Liquid limit <strong>an</strong>d linear shrinkage 191<br />

10.4 Free swell <strong>an</strong>d other index properties 194<br />

10.5 Clay fraction <strong>an</strong>d linear shrinkage, free swell 198<br />

Chapter 11. Conclusions 200<br />

Chapter 12. Recommendations 207<br />

References<br />

Appendices<br />

Appendix A. Results <strong>of</strong> oedometer consolidation tests<br />

Appendix B. Results <strong>of</strong> swelling tests on black <strong>clays</strong><br />

Appendix C. Results <strong>of</strong> swelling tests on black <strong>clays</strong><br />

Appendix D. Distribution/ variation <strong>of</strong> index <strong>an</strong>d <strong>engineering</strong> properties <strong>of</strong><br />

black <strong>clays</strong> in Nairobi area<br />

Appendix E. Geotechnical soil map <strong>of</strong> Nairobi area<br />

Figures<br />

1.1 Location map <strong>of</strong> Nairobi city <strong>an</strong>d study area 3<br />

3.1 Geology: relative position <strong>of</strong> Great Rift Valley, extent <strong>of</strong> early lava flows <strong>an</strong>d<br />

ages <strong>of</strong> rock formations in Nairobi area 22-23<br />

4.1 Site investigation/ sampling scheme employed for black <strong>clays</strong> in this study 31<br />

4.2 Field depth sounding results for black <strong>clays</strong> <strong>an</strong>d red soils 35<br />

4.3 Contoured depth variations in black <strong>clays</strong> <strong>an</strong>d red soils 36<br />

4.4 Block diagrams <strong>of</strong> depth variations in black <strong>clays</strong> <strong>an</strong>d red soils 37


iv<br />

4.5 Principle <strong>of</strong> v<strong>an</strong>e shear test 38<br />

4.6 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with soil depth 41<br />

4.7 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with natural moisture content <strong>of</strong> soils 42<br />

4.8 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with bulk density <strong>of</strong> soils 43<br />

5.1 Distribution <strong>of</strong> soils <strong>of</strong> the study area 47<br />

6.1 – 6.6 X-ray diffraction diagrams <strong>of</strong> black <strong>clays</strong> 61-64<br />

6.7 – 6.10 X-ray diffraction diagrams <strong>of</strong> red soils 65-67<br />

6.11 Variation <strong>of</strong> carbon content <strong>of</strong> red soils with depth 80<br />

6.12 Variation <strong>of</strong> carbon content <strong>of</strong> black <strong>clays</strong> with depth 81<br />

6.13 Variation <strong>of</strong> carbon content <strong>of</strong> both black <strong>clays</strong> <strong>an</strong>d red soils with depth 82<br />

7.1 A plasticity chart classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 90<br />

7.2 Activity level classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 91<br />

7.3a Particle size distribution curves for black <strong>clays</strong> <strong>an</strong>d red soils 102<br />

7.3b Tri<strong>an</strong>gular classification chart for black <strong>clays</strong> <strong>an</strong>d red soils 103<br />

7.4 Shear stress/ displacement curves <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 107<br />

7.5 Coulomb or shear failure envelopes obtained for black <strong>clays</strong> <strong>an</strong>d red soils 110<br />

7.6 Correlation between shear paramters (c´, φ´) <strong>an</strong>d relative consistency,<br />

Cr, as well as moisture content, Wn, <strong>of</strong> black <strong>clays</strong> 111-112<br />

7.7 Correlation between shear paramters (c´, φ´) <strong>an</strong>d plasticity index (PI)<br />

<strong>of</strong> black <strong>clays</strong> 112-113<br />

7.8 Correlation between shear paramters (c´, φ´) <strong>an</strong>d reciprocal <strong>of</strong><br />

plasticity index (1/PI) <strong>of</strong> black <strong>clays</strong> 113<br />

7.9. Correlation between shear paramters (c´, φ´) <strong>an</strong>d ratio Cr/PI<br />

for black <strong>clays</strong> 114<br />

7.10 – 7.12 Time factor Tv related to degree <strong>of</strong> consolidation U% 117-118<br />

7.13 Log-time/settlement curve <strong>an</strong>alysis method showing phases<br />

<strong>of</strong> consolidation 118<br />

7.14 Analysis <strong>of</strong> square-root time/ settlement curve 118<br />

7.15a. Log-pressure/ voids ratio (e/ log p) curve 122<br />

7.15b. Determination <strong>of</strong> coefficient <strong>of</strong> secondary compression cα 122<br />

7.16 Conventional log-time/settlement curves for black <strong>clays</strong> during<br />

loading <strong>an</strong>d unloading stages 127<br />

7.17 Unconventional log-time/settlement curves for red soils<br />

during loading <strong>an</strong>d unloading stages 127<br />

7.18 Voids ratio/ log Pressure (e/ log p) curves for black <strong>clays</strong> <strong>an</strong>d red soils 128<br />

7.19 - 7.20 Permeability/ log pressure (K/log p) curves for black <strong>clays</strong><br />

<strong>an</strong>d red soils 130-131<br />

7.21 Coefficient <strong>of</strong> volume compressibility/ log pressure (mv/log p)<br />

curves for black <strong>clays</strong> <strong>an</strong>d red soils 131<br />

7.22 Coefficient <strong>of</strong> consolidation/ log pressure (cv/log p) curves for black<br />

<strong>clays</strong> <strong>an</strong>d red soils 132<br />

7.23 – 7.24 Correlation between laboratory determined liquid limits <strong>an</strong>d<br />

compression indices for black <strong>clays</strong> <strong>an</strong>d red soils 134<br />

7.25 – 7.26 Correlation between swell index <strong>an</strong>d Atterberg<br />

limits for black <strong>clays</strong> <strong>an</strong>d red soils 137<br />

7.27 Correlation between swell index <strong>an</strong>d free swell for black <strong>clays</strong><br />

<strong>an</strong>d red soils 137<br />

7.28 Swelling pressure curves obtained for black <strong>clays</strong> 139


v<br />

7.29 – 7.33 Correlation between swelling pressure <strong>an</strong>d index properties<br />

<strong>of</strong> black <strong>clays</strong> 140- 141<br />

7.34 Variation <strong>of</strong> percentage swelling with loading pressure in black <strong>clays</strong> 143<br />

7.35 Correlation; % swelling <strong>an</strong>d loading pressure <strong>of</strong> black clay samples 144<br />

7.36 Correlation between % swelling <strong>an</strong>d loading pressure <strong>of</strong> black <strong>clays</strong> 146<br />

7.37 Correlation; percentage swelling in relation to initial specimen height, Ho,<br />

<strong>an</strong>d extent <strong>of</strong> external loading for black <strong>clays</strong> 149<br />

7.38 Estimation <strong>of</strong> swelling properties <strong>of</strong> black <strong>clays</strong> using Greek method 151<br />

8.1-8.2 Natural moisture content variation in black <strong>clays</strong> across the study area for<br />

depths less th<strong>an</strong> 0,50m as well as those greater th<strong>an</strong> 0,50m 155-157<br />

8.3-8.4 Liquid limit variation in black <strong>clays</strong> across the study area for depths <strong>of</strong><br />

less th<strong>an</strong> 0,50m as well as those greater th<strong>an</strong> 0,50m 158-161<br />

183<br />

8.5-8.6 Plasticity index variation in black <strong>clays</strong> across the study area for the depth <strong>of</strong><br />

less th<strong>an</strong> 0,5m as well as that greater th<strong>an</strong> 0,5m 162-163<br />

8.7-8.8 Linear shrinkage variation across the study area for the depth <strong>of</strong><br />

less th<strong>an</strong> 0,50m as well as that greater th<strong>an</strong> 0,50m 164-165<br />

8.9-8.10 Free swell variation in black <strong>clays</strong> across the study area at depths<br />

<strong>of</strong> less th<strong>an</strong> 0,5m as well as those greater th<strong>an</strong> 0,50m 166-168<br />

8.11 Variation <strong>of</strong> clay proportion in black <strong>clays</strong> across the study area at depths <strong>of</strong> less<br />

th<strong>an</strong> 0,5m as well as those greater th<strong>an</strong> 0,50m 170<br />

8.12-8.13 Variation <strong>of</strong> fines proportion in black <strong>clays</strong> across the study area at depths<br />

<strong>of</strong> less th<strong>an</strong> 0,5m as well as those greater th<strong>an</strong> 0,50m 171-172<br />

8.14 Distribution <strong>an</strong>d variation <strong>of</strong> coarse fraction in black <strong>clays</strong> 173<br />

8.15 Distribution <strong>an</strong>d variation <strong>of</strong> shear strength parameters (c´, φ´) in black<br />

<strong>clays</strong> 174-175<br />

10.1 Correlation between laboratory measured values <strong>of</strong> plasticity index <strong>an</strong>d linear<br />

shrinkage <strong>of</strong> black <strong>clays</strong><br />

10.2 Correlation between calculated <strong>an</strong>d laboratory measured plasticity indices<br />

<strong>of</strong> black <strong>clays</strong> 184<br />

10.3 Correlation between laboratory measured values <strong>of</strong> plasticity index <strong>an</strong>d linear<br />

shrinkage <strong>of</strong> red soils 186<br />

10.4 Correlation between calculated <strong>an</strong>d laboratory measured plasticity indices<br />

<strong>of</strong> red soils 187<br />

10.5 Correlation between laboratory measured values <strong>of</strong> plasticity index<br />

<strong>an</strong>d liquid limit <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 187<br />

10.6 Correlation between calculated <strong>an</strong>d laboratory determined plasticity indices<br />

<strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 191<br />

10.7 Correlation: laboratory measured values <strong>of</strong> linear shrinkage <strong>an</strong>d liquid<br />

limit <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 191<br />

10.8 A correlation between calculated <strong>an</strong>d laboratory determined linear<br />

shrinkage <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 194<br />

10.9 Correlation: laboratory measured values <strong>of</strong> free swell <strong>an</strong>d those <strong>of</strong> liquid<br />

limit, plasticity index <strong>an</strong>d linear shrinkage <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 197<br />

10.10 Correlation between calculated <strong>an</strong>d laboratory determined free<br />

swell values <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 198<br />

10.11 Correlation between laboratory measured values <strong>of</strong> linear shrinkage <strong>an</strong>d<br />

clay fraction <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 198<br />

10.12 Correlation between laboratory measured values <strong>of</strong> free swell <strong>an</strong>d clay<br />

fraction <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 199


vi<br />

Tables<br />

1.1 Me<strong>an</strong> maximum <strong>an</strong>d minimum temperatures for each month <strong>of</strong> the<br />

year over a period <strong>of</strong> 50 years, for Nairobi city <strong>an</strong>d the present study area 11<br />

1.2 Average rainfall, me<strong>an</strong> relative humidity <strong>an</strong>d amount <strong>of</strong> bright sunshine<br />

for each month <strong>of</strong> the year, for Nairobi city <strong>an</strong>d the present study area 12<br />

2.1 Stratigraphic correlation <strong>of</strong> rocks in the Nairobi region 17<br />

3.1 Chemical <strong>an</strong>alyses <strong>of</strong> alkali-feldspars from the Kapiti phonolite<br />

<strong>an</strong>d similar rocks 25<br />

3.2 Chemical <strong>an</strong>alyses <strong>of</strong> strongly alkaline lavas <strong>of</strong> Nairobi phonolite<br />

<strong>an</strong>d other rocks 27<br />

3.3 Chemical <strong>an</strong>alyses <strong>of</strong> mildly alkaline lavas <strong>of</strong> Nairobi trachyte<br />

<strong>an</strong>d other volc<strong>an</strong>ic rocks 28<br />

3.4 Mineral compositions <strong>of</strong> trachytes <strong>an</strong>d phonolites from Nairobi <strong>an</strong>d<br />

the study area 29<br />

4.1 Classification <strong>of</strong> s<strong>of</strong>t soils based on v<strong>an</strong>e shear strength 40<br />

4.2 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t black <strong>clays</strong> at Madaraka West 40<br />

4.3 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t red soils at Arboretum 40<br />

4.4 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t red soils at Kenya High School 41<br />

5.1 Pr<strong>of</strong>ile description <strong>of</strong> a red friable clay 49<br />

5.2 Soluble base content in soils <strong>of</strong> the study area 49<br />

5.3 Pr<strong>of</strong>ile description <strong>of</strong> a shallow yellow-brown clay overlying laterite 50<br />

5.4 Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> red soils obtained in this study 51<br />

5.5 Earlier results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> <strong>clays</strong> <strong>an</strong>d red soil 52<br />

5.6 Some results <strong>of</strong> soil classification tests on red soils in present study 52<br />

5.7 Earlier results <strong>of</strong> soil classification tests on red <strong>clays</strong> from the Nairobi area 53<br />

5.8 Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> black <strong>clays</strong> obtained in this study 54<br />

5.9 Clay mineral composition <strong>of</strong> black <strong>clays</strong> 56<br />

5.10 Pr<strong>of</strong>ile description <strong>of</strong> black clay 57<br />

6.1 Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> red soils obtained in this study 58<br />

6.2 Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> black <strong>clays</strong> obtained in this study 59<br />

6.3 Clay mineral composition <strong>of</strong> black <strong>clays</strong> 61<br />

6.4 Estimated mineralogical composition <strong>of</strong> red soils 64<br />

6.5 Theoretical compositions <strong>of</strong> kaolinite, illite <strong>an</strong>d montmorillonite 65<br />

6.6. Carbon content <strong>of</strong> soils <strong>of</strong> the study area 80<br />

7.1 Activity <strong>of</strong> clay soils <strong>an</strong>d clay minerals 85<br />

7.2 Results <strong>of</strong> index tests performed on black <strong>clays</strong> <strong>an</strong>d red soils 87-88<br />

7.3 Atterberg limits, activity/ free swell classification <strong>an</strong>d assessment<br />

<strong>of</strong> possible clay mineralogy <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 92<br />

7.4 Free swell classification <strong>of</strong> clay soils (Gibbs & Holtz, 1956) 94<br />

7.5 Soil classification by particle size (BS 1377: 1975) 97<br />

7.6 Viscosity <strong>an</strong>d density <strong>of</strong> water (Kaye <strong>an</strong>d Laby, 1973) 99<br />

7.7 Results <strong>of</strong> grain size <strong>an</strong>alysis for black <strong>clays</strong> <strong>an</strong>d red soils 100-101<br />

7.8 Typical values <strong>of</strong> φ for dry noncohesive soils <strong>an</strong>d clay (after Lambe <strong>an</strong>d<br />

Whitm<strong>an</strong>, 1979) 106<br />

7.9 Shear strength parameters <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 108<br />

7.10 Distribution <strong>an</strong>d/variation <strong>of</strong> shear strength parameter (φ´) across study area 109<br />

7.11 Time factors for one-dimensional consolidation (Leonards, 1962) 117<br />

7.12 Derived parameters from results <strong>of</strong> oedometer consolidation tests<br />

for the loading r<strong>an</strong>ge <strong>of</strong> 25-800 kPa 128


vii<br />

136<br />

139<br />

7.13 Compressibility classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 129<br />

7.14 Rates <strong>of</strong> consolidation-settlement <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils 130<br />

7.15a Compression indices, Cc, <strong>an</strong>d compressibility classification <strong>of</strong> clay soils 132<br />

7.15b Results <strong>of</strong> compression indices correlated with Atterberg limits 133<br />

7.16 Correlation between calculated <strong>an</strong>d laboratory derived compression indices 135<br />

7.17 Classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils based on coefficient <strong>of</strong> secondary<br />

compression 136<br />

7.18 Results <strong>of</strong> swell indices for black <strong>clays</strong> <strong>an</strong>d red soils correlated with Atterberg<br />

limits <strong>an</strong>d free swell<br />

7.19 Results <strong>of</strong> swelling pressures for black <strong>clays</strong>, alongside corresponding<br />

index properties <strong>an</strong>d clay mineralogy<br />

7.20 Percentage swelling <strong>of</strong> black <strong>clays</strong> under various load decrements 142<br />

7.21 Relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> under various loads 143<br />

7.22 Classification <strong>of</strong> relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> based on<br />

percentage swelling 144<br />

7.23 Percentage swelling <strong>of</strong> black <strong>clays</strong> under various load decrements (S% is in<br />

relation to initial specimen height, Ho, i.e. 2 nd method in this study) 147<br />

7.24 Classification <strong>of</strong> relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> using the<br />

newly derived/ 2 nd method 147<br />

7.25 Computed values for estimating swelling characteristics <strong>of</strong> black <strong>clays</strong><br />

using the Greek method 150<br />

8.1 Values <strong>of</strong> index properties <strong>of</strong> soils along field pr<strong>of</strong>iles in this study 156<br />

8.2 Values <strong>of</strong> clay <strong>an</strong>d fines fractions <strong>of</strong> soils along field pr<strong>of</strong>iles in this study 169<br />

10.1 Calculated <strong>an</strong>d laboratory measured values <strong>of</strong> plasticity index <strong>of</strong> black <strong>clays</strong> 181<br />

10.2 Calculated <strong>an</strong>d laboratory measured values <strong>of</strong> plasticity index <strong>of</strong> red soils 182<br />

10.3 Calculated <strong>an</strong>d laboratory measured plasticity indices <strong>of</strong> black <strong>clays</strong> 188<br />

10.4 Calculated <strong>an</strong>d laboratory measured plasticity indices <strong>of</strong> red soils 190<br />

10.5 Calculated <strong>an</strong>d laboratory measured linear shrinkage values <strong>of</strong> black <strong>clays</strong><br />

<strong>an</strong>d red soils 192<br />

10.6 Calculated <strong>an</strong>d laboratory measured free swell values <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d<br />

red soils 195<br />

Plates<br />

1.1 (a) & (b) New stone structures coming up in the study area 4<br />

1.2 (a) & (b) Strong shrinkage cracks in black cotton soils during dry months 5<br />

1.3 (a) & (b) Cracked <strong>an</strong>d peeled-<strong>of</strong>f asphalt on tarmac road giving way<br />

to formation <strong>of</strong> pot holes in a zone <strong>of</strong> black <strong>clays</strong> in the Nairobi city centre 5<br />

1.4 A slightly tilting building in a zone <strong>of</strong> black <strong>clays</strong> in the study area 6<br />

1.5 (a) & (b) Light/ low rise buildings located in black <strong>clays</strong> with poor drainage 6<br />

1.6 (a) & (b) Heavy constructed structures in a zone <strong>of</strong> black <strong>clays</strong> in<br />

the study area 7<br />

1.7 (a) & (b) View <strong>of</strong> the Great Rift Valley from its eastern fl<strong>an</strong>k, to the west <strong>of</strong><br />

Nairobi area 8<br />

1.8 (a) & (b) Flat plains characterising areas <strong>of</strong> black cotton soils in this study 9<br />

4.1 (a) & (b) H<strong>an</strong>d-dug excavation pits <strong>an</strong>d undisturbed sampling in black <strong>clays</strong> 32<br />

4.2 (a) & (b) H<strong>an</strong>d-dug excavation pits <strong>an</strong>d undisturbed sampling in red soils 33


viii<br />

4.3 Field augering <strong>of</strong> soils 33<br />

4.4 Cone penetration sounding <strong>of</strong> soil depths 33<br />

4.5 (a) & (b) V<strong>an</strong>e shear testing in s<strong>of</strong>t black <strong>clays</strong> 39<br />

6.1 – 6.5 Results <strong>of</strong> SEM <strong>an</strong>alysis <strong>of</strong> Nairobi phonolite 68-69<br />

6.6 – 6.9 Results <strong>of</strong> SEM <strong>an</strong>alysis <strong>of</strong> Nairobi trachyte 70<br />

6.10 - 6.24 Results <strong>of</strong> SEM <strong>an</strong>alysis <strong>of</strong> black <strong>clays</strong> 71-75<br />

6.25 – 6.31 Results <strong>of</strong> SEM <strong>an</strong>alysis <strong>of</strong> red soils 75-77<br />

7.1 Plastic limit determination <strong>of</strong> clay soils 86<br />

7.2 Liquid limit determination <strong>of</strong> clay soils 89<br />

7.3 Free swell testing <strong>of</strong> clay soils 94<br />

7.4 Linear shrinkage testing <strong>of</strong> clay soils 96<br />

7.5 Sieves used for grain size <strong>an</strong>alysis <strong>of</strong> soils 97<br />

7.6 Sedimentation procedure in grain size <strong>an</strong>alysis <strong>of</strong> soils 97<br />

7.7 Direct shear testing <strong>of</strong> clay soils 105<br />

7.8 Oedometer consolidation testing <strong>of</strong> clay soils 125<br />

7.9 Arr<strong>an</strong>gement for swelling pressure determination <strong>of</strong> black <strong>clays</strong> 138


ix<br />

Acknowledgements<br />

I would like to th<strong>an</strong>k all those who contributed in different ways towards making my Ph.D.<br />

research work a success.<br />

On the Germ<strong>an</strong> side, my gratitude goes to the following: the Germ<strong>an</strong> Katholic Academic<br />

Exch<strong>an</strong>ge Programme (KAAD), in Bonn, for funding my studies through a four-year Ph.D.<br />

scholarship; staff <strong>of</strong> the Afric<strong>an</strong> Department, KAAD, Bonn (Dr. T. Scheidtweiler <strong>an</strong>d MS.<br />

Simone Saure) for valuable general information, advice <strong>an</strong>d guid<strong>an</strong>ce throughout the<br />

scholarship period; the Institute <strong>of</strong> Geology <strong>an</strong>d Palaeontology, Technical Universty <strong>of</strong><br />

Clausthal, for availing facilities <strong>an</strong>d equipment for field <strong>an</strong>d laboratory investigations; the<br />

Institute <strong>of</strong> Mineralogy (through Pr<strong>of</strong>. K. Mengel), Technical Universty <strong>of</strong> Clausthal, for<br />

facilitating chemical <strong>an</strong>d mineralogical <strong>an</strong>alyses <strong>of</strong> soil <strong>an</strong>d rock materials; Pr<strong>of</strong>. G. Reik<br />

(Institute <strong>of</strong> Geology <strong>an</strong>d Palaeontology, TU-Clausthal) for supervision <strong>of</strong> field <strong>an</strong>d<br />

laboratory studies as well as reading <strong>an</strong>d correcting the thesis; Pr<strong>of</strong>. H. J. Gursky (Institute <strong>of</strong><br />

Geology <strong>an</strong>d Palaeontology, TU-Clausthal) for reading <strong>an</strong>d correcting the thesis; <strong>an</strong>d the<br />

Faculty Examination Board under Pr<strong>of</strong>. D. Mayer (Faculty <strong>of</strong> Mathematics <strong>an</strong>d Natural<br />

Sciences, TU-Clausthal), for examining the thesis work. I also appreciate the assist<strong>an</strong>ce<br />

rendered by other members <strong>of</strong> staff, research assist<strong>an</strong>ts <strong>an</strong>d students <strong>of</strong> the Institute <strong>of</strong><br />

Geology <strong>an</strong>d Palaeontology, TU-Clausthal, in areas <strong>of</strong> samples preparation <strong>an</strong>d <strong>an</strong>alysis, data<br />

<strong>an</strong>alysis <strong>an</strong>d interpretation, as well as printing <strong>of</strong> the thesis.<br />

On the Keny<strong>an</strong> side, my pr<strong>of</strong>ound gratitude goes to the following: the Ministry <strong>of</strong> Education<br />

for gr<strong>an</strong>ting me a research permit to conduct fieldwork in Nairobi; the Kenya Airports<br />

Authority (KAA) for permitting field investigations around JKIA <strong>an</strong>d Wilson airports; the<br />

Survey <strong>of</strong> Kenya for availing topographic maps; Department <strong>of</strong> Geology, University <strong>of</strong><br />

Nairobi, for storage <strong>of</strong> field equipment; Dr. Patrick L. Legge (Department <strong>of</strong> Geology,<br />

University <strong>of</strong> Nairobi,) for assisting in the field study phase; Mines <strong>an</strong>d Geology Department,<br />

Nairobi, for permitting shipment <strong>of</strong> soil <strong>an</strong>d rock samples to Germ<strong>an</strong>y for laboratory <strong>an</strong>alysis.<br />

I also th<strong>an</strong>k Dr. Chris Nyamai (Department <strong>of</strong> Geology, University <strong>of</strong> Nairobi) as well as the<br />

field investigation <strong>an</strong>d sampling crew for their contributions.<br />

I would also like to th<strong>an</strong>k my wife (Ragchaasuren S. ) <strong>an</strong>d parents, for their tireless support,<br />

encouragement <strong>an</strong>d patience. Contributions <strong>an</strong>d assist<strong>an</strong>ce rendered by friends in Clausthal,<br />

Gaston Schettkatt <strong>an</strong>d Nassim Ouassa, are highly appreciated.


x<br />

Summary (Abstract)<br />

An <strong>engineering</strong> <strong>geological</strong> <strong>characterisation</strong> <strong>of</strong> <strong>tropical</strong> <strong>clays</strong>. Case study: Clay soils <strong>of</strong> Nairobi, Kenya<br />

Simon W<strong>an</strong>yonyi Ot<strong>an</strong>do, M.Sc.<br />

The purpose <strong>of</strong> this project was to study <strong>an</strong>d evaluate soils <strong>engineering</strong> <strong>geological</strong> characteristics with a view to determining their form <strong>of</strong> distribution <strong>an</strong>d<br />

spatial variation across the study area. It also aimed at assessing the suitability, capacity <strong>an</strong>d/ or limitations <strong>of</strong> the soils to support low/ light constructed<br />

<strong>engineering</strong> structures, as well as recommending best suited <strong>an</strong>d economical soil improvement/stabilisation techniques where necessary. The study <strong>an</strong>d<br />

investigation was therefore centered on determining/ developing statistical models to describe <strong>an</strong>d predict the distribution <strong>of</strong> soils <strong>engineering</strong> properties.<br />

Results <strong>of</strong> statistical <strong>an</strong>alyses served to determine the degree <strong>of</strong> homogeneity or uniformity <strong>of</strong> the soils. It also aimed at developing/ preparing small-scale<br />

<strong>an</strong>d site specific <strong>engineering</strong> <strong>geological</strong> maps to provide spatially continouos ground information, soil data <strong>an</strong>d depth variations across the study area. The<br />

study also sought to present results <strong>of</strong> soil classifications <strong>an</strong>d correlations between soil index <strong>an</strong>d/ or <strong>engineering</strong> properties. A detailed chemical/<br />

mineralogical <strong>an</strong>alysis <strong>an</strong>d evaluation <strong>of</strong> the soils <strong>an</strong>d underlying bedrock was also carried out to investigate <strong>an</strong>d establish the genesis <strong>of</strong> the soils.<br />

The current study area lies mainly to the south <strong>of</strong> the fast growing Nairobi city, Kenya. It is covered by topographic sheet 148/4 <strong>of</strong> Nairobi from Survey <strong>of</strong><br />

Kenya, bounded by latitudes 1°15´S <strong>an</strong>d 1°24´S to the north <strong>an</strong>d south; <strong>an</strong>d longitudes 36°45´E <strong>an</strong>d 36°58´E to the west <strong>an</strong>d east ; respectively, <strong>an</strong>d<br />

occupies <strong>an</strong> area <strong>of</strong> about 367 km². The area is covered mainly by varying thickness <strong>of</strong> black cotton soils/ black <strong>clays</strong>. Red soils are limited to the northwestern<br />

part <strong>of</strong> the area. Underlying the soils are Tertiary volc<strong>an</strong>ics which inturn overly a Basement system <strong>of</strong> folded Precambri<strong>an</strong> metamorphic rocks. The<br />

research <strong>an</strong>d study was centered on black <strong>clays</strong> with red soils included for comparison purposes.<br />

The present area was chosen because it is mainly located on the outskirts <strong>of</strong> Nairobi city which are projected for future exp<strong>an</strong>sion <strong>of</strong> the town, as regards<br />

construction <strong>of</strong> new social-economic <strong>an</strong>d industrial/residential facilities. There is also a shiftimg trend in building tradition from timber <strong>an</strong>d iron-sheet<br />

structures on strip/ pad footings towards masonry construction <strong>of</strong> low buildings on slab foundations using dimension stone <strong>an</strong>d bricks, hence the need for <strong>an</strong><br />

<strong>engineering</strong> properties <strong>characterisation</strong> <strong>of</strong> the soils. Current field observations also show the black <strong>clays</strong> to be potentially exp<strong>an</strong>sive/ reactive <strong>an</strong>d exhibit<br />

ground instabilities, failures <strong>an</strong>d movements; with serious environmental, <strong>engineering</strong> <strong>an</strong>d social-economic implications.<br />

The study incorporated suitable field procedures, sampling criteria <strong>an</strong>d laboratory chemical/ mineralogical <strong>an</strong>d <strong>engineering</strong> soils testing methods. The field<br />

study methods included cone penetration sounding, soil augering, in-situ description <strong>of</strong> soils <strong>an</strong>d underlying rock, tests for possible carbonate content, as<br />

well as undisturbed <strong>an</strong>d disturbed soil sampling. This study phase had the objectives <strong>of</strong> providing information on the nature <strong>an</strong>d distribution <strong>of</strong> exp<strong>an</strong>sive/<br />

reactive black cotton soils as well as red soils <strong>of</strong> the study area. It also served to acquire site specific information on environmental characteristics<br />

(groundwater conditions, climate, vegetation) which impact on in situ soil behaviour, investigate <strong>an</strong>d model depth variation <strong>of</strong> soils; as well as compare in<br />

situ soil properties/behaviour <strong>an</strong>d derived parameters with those predicted <strong>an</strong>d/ or derived from laboratory studies.<br />

Laboratory studies performed on soil samples collected from the field included index tests <strong>of</strong> natural moisture content, Atterberg/plasticity limits, swelling<br />

capability/free swell <strong>an</strong>d linear shrinkage, as well as grain size <strong>an</strong>alysis. Direct shear strength tests, Oedometer consolidation/ compressibilty tests,<br />

percentage swelling <strong>an</strong>d swelling pressure tests were also carried out by performing the tests on undisturbed samples. Laboratory chemical/ mineralogical<br />

<strong>an</strong>alyses <strong>an</strong>d/ or studies employed the use <strong>of</strong> X-ray fluorescence (XRF), X-ray diffraction (XRD) <strong>an</strong>d sc<strong>an</strong>ning electron microscope (SEM) techniques.<br />

The field <strong>an</strong>d laboratory investigations/ studies <strong>of</strong> soils <strong>of</strong> the project area have provided a r<strong>an</strong>ge <strong>of</strong> soil index <strong>an</strong>d <strong>engineering</strong> properties as well as<br />

chemical <strong>an</strong>d mineralogical composition. The results <strong>of</strong> study show the distribution <strong>an</strong>d spatial variation <strong>of</strong> index properties <strong>an</strong>d chemical/ mineralogical<br />

composition <strong>of</strong> black <strong>clays</strong> to be generally uniform across the study area, implying a generally homogeneous occurrence <strong>an</strong>d <strong>engineering</strong> character <strong>of</strong> the<br />

<strong>clays</strong>. Results <strong>of</strong> chemical/mineralogical <strong>an</strong>alyses show the black <strong>clays</strong> to be composed mainly <strong>of</strong> smectites (montmorillonite) as well as having been<br />

largely genetically derived from the gradual conversion <strong>of</strong> lacustrine deposits <strong>of</strong> volc<strong>an</strong>ic ash, colluvium, alluvium <strong>an</strong>d s<strong>of</strong>t Pleistocene materials previously<br />

deposited in a basin-like lake <strong>an</strong>d swampy environment during a pluvial period. Contributions by leached components from areas <strong>of</strong> red soils as well as<br />

eroded materials from surrounding areas are to a lesser extent.<br />

Index properties <strong>of</strong> red soils point to a comparatively more stable <strong>an</strong>d reliable <strong>engineering</strong> character <strong>of</strong> the soils. Results <strong>of</strong> chemical/ mineralogical <strong>an</strong>alyses<br />

show these soils to be mainly kaolinite in composition <strong>an</strong>d polygenetic in origin. The soils must have derived/developed <strong>an</strong>d formed under humid conditions<br />

by weathering/alteration <strong>of</strong> underlying Nairobi trachytes/ volc<strong>an</strong>ic tuff <strong>an</strong>d volc<strong>an</strong>ic ash. A limited contribution was derived from eroded <strong>an</strong>d watertr<strong>an</strong>sported<br />

detrital materials.<br />

The index properties <strong>an</strong>d grain sizes have been used to classify black <strong>clays</strong> into very high to extremely high plasticity <strong>clays</strong> <strong>an</strong>d/ or silty <strong>clays</strong> exhibiting<br />

medium to very high levels <strong>of</strong> activity as well as high swelling capabilities/ potential exp<strong>an</strong>siveness on wetting from a dry condition. This <strong>engineering</strong><br />

behaviour could be accounted for by the high content <strong>of</strong> the more exp<strong>an</strong>sive clay minerals <strong>of</strong> smectites (90% <strong>an</strong>d over) in the black <strong>clays</strong>. Likewise, the red<br />

soils have been classified as medium to high plasticity clayey silts <strong>an</strong>d silty <strong>clays</strong> with s<strong>an</strong>d showing medium activity levels <strong>an</strong>d generally low swelling<br />

capabilities/ potential exp<strong>an</strong>siveness when allowed free access to water. These characteristics have been attributed to the high content (80% <strong>an</strong>d over) <strong>of</strong> the<br />

rather less exp<strong>an</strong>sive clay mineral, kaolinite, in the red soils.<br />

Results <strong>of</strong> laboratory shear strength investigations showed the black <strong>clays</strong> to be limited in strength <strong>an</strong>d have relatively low <strong>an</strong>gles <strong>of</strong> shear resist<strong>an</strong>ce<br />

averaging about 18°. The red soils were found to be relatively more stable, with <strong>an</strong>gles <strong>of</strong> shear resist<strong>an</strong>ce <strong>of</strong> 28° to 29°. On the other h<strong>an</strong>d, results <strong>of</strong><br />

laboratory oedometer consolidation tests show the red soils to exhibit medium to very high compressibility <strong>an</strong>d medium to high rates <strong>of</strong> consolidationsettlement<br />

on normal external loading. This could be attributed to their loose friable nature <strong>an</strong>d relatively higher porosity <strong>an</strong>d permeability which facilitate<br />

faster drainage on external loading. These soils would therefore be expected to undergo rapid consolidation – settlements, especially during the construction<br />

stage, without posing long-term instability problems to constructed structures. However, the black <strong>clays</strong> would exhibit slightly lower compressibility <strong>an</strong>d<br />

low rates <strong>of</strong> consolidation-settlement due to their compact/cohesive nature as well as relatively lower porosity <strong>an</strong>d permeability which tend to limit<br />

drainage. In practice therefore, settlement <strong>an</strong>d instability <strong>of</strong> structures located on these <strong>clays</strong> would be expected to persist beyond the construction stage.<br />

Results <strong>of</strong> statistical <strong>an</strong>alysis <strong>an</strong>d correlations show that most soil index properties could be characterised <strong>an</strong>d/ or estimated with a high degree <strong>of</strong><br />

approximation from results <strong>of</strong> Atterberg limits <strong>an</strong>d other index tests performed on disturbed <strong>an</strong>d fractioned samples. New relationships have also been<br />

derived <strong>an</strong>d developed to estimate <strong>an</strong>d predict percentage swelling <strong>an</strong>d swelling pressure characteristics <strong>of</strong> black <strong>clays</strong> in situ under various loading<br />

conditions. However, in situ <strong>characterisation</strong> <strong>of</strong> <strong>engineering</strong> behaviour <strong>of</strong> soils as regards shear strength, consolidation-settlement, swelling pressure <strong>an</strong>d<br />

percentage swelling was shown to be better accomplished by direct measurements on undisturbed core samples rather th<strong>an</strong> through correlation with results<br />

<strong>of</strong> index tests performed on disturbed/ part samples which harbour destroyed soil fabric <strong>an</strong>d structure.<br />

The vast data base <strong>of</strong> soil index <strong>an</strong>d <strong>engineering</strong> properties as well as chemical/ mineralogical composition obtained from field <strong>an</strong>d laboratory<br />

investigations/ studies <strong>of</strong> soils <strong>of</strong> the project area, together with prepared <strong>engineering</strong> <strong>geological</strong>/ geotechnical maps would be readily useful <strong>an</strong>d applicable<br />

by <strong>engineering</strong> geologists, geo-engineers <strong>an</strong>d pl<strong>an</strong>ners <strong>of</strong> future-intended construction projects in the area. The results <strong>of</strong> the investigations together with<br />

related <strong>engineering</strong> <strong>geological</strong> assessments would assist especially at the stages <strong>of</strong> pl<strong>an</strong>ning, designing <strong>an</strong>d construction <strong>of</strong> projected facilities; <strong>an</strong>d also<br />

subsequent mainten<strong>an</strong>ce after completion. The results will also serve to complement on-going <strong>an</strong>d future research on the <strong>characterisation</strong> <strong>of</strong> exp<strong>an</strong>sive <strong>an</strong>d<br />

reactive black cotton soils, especially as regards their negative environmental, <strong>engineering</strong> <strong>an</strong>d social-economic implications.


1<br />

Chapter 1<br />

Introduction<br />

1.1 Scope <strong>of</strong> study<br />

The purpose <strong>of</strong> this project was to study <strong>an</strong>d evaluate <strong>engineering</strong> <strong>geological</strong> characteristics <strong>of</strong> soils<br />

found in the Nairobi area, Kenya. The information <strong>an</strong>d data so obtained would be useful for<br />

assessing the suitability <strong>an</strong>d capacity <strong>of</strong> soils found regarding the possibilities <strong>an</strong>d / or limitations <strong>of</strong><br />

developing <strong>an</strong>d constructing low buildings <strong>an</strong>d other light <strong>engineering</strong> structures ( roads, pavements<br />

etc) in the study area <strong>an</strong>d <strong>geological</strong>ly similar neighbourhoods. Attempts have been made by the<br />

author to present the <strong>geological</strong> data <strong>an</strong>d soil parameters in a spatial continuous form that would be<br />

readily useful <strong>an</strong>d applicable by <strong>engineering</strong> geologists, geo-engineers <strong>an</strong>d pl<strong>an</strong>ners <strong>of</strong> futureintended<br />

construction projects in the area. The results <strong>of</strong> the investigations together with related<br />

<strong>engineering</strong> <strong>geological</strong> assessments would assist especially at the stages <strong>of</strong> pl<strong>an</strong>ning, designing <strong>an</strong>d<br />

construction <strong>of</strong> projected facilities; <strong>an</strong>d also subsequent mainten<strong>an</strong>ce after completion.<br />

The current study area lies mainly to the south <strong>of</strong> the fast growing Nairobi city, Kenya. However,<br />

the southern half <strong>of</strong> the Nairobi metropolit<strong>an</strong>, including the City centre, is also covered. The area is<br />

covered by topographic sheet 148/4 <strong>of</strong> Nairobi, at a scale <strong>of</strong> 1:50 000; <strong>an</strong>d printed by the Survey <strong>of</strong><br />

Kenya (1991). The exact location <strong>an</strong>d extent <strong>of</strong> the study area is described in the next section. The<br />

area is covered mainly by varying thickness <strong>of</strong> black cotton soils. Occurrences <strong>of</strong> red soils are also<br />

found, but limited in the north-western part <strong>of</strong> the area. Underlying the soils are Tertiary volc<strong>an</strong>ics<br />

which in turn overly a Basement system <strong>of</strong> folded Precambri<strong>an</strong> metamorphic rocks.<br />

The present area was chosen because it is mainly located on the outskirts <strong>of</strong> Nairobi city which are<br />

projected for future exp<strong>an</strong>sion <strong>of</strong> the town, as regards construction <strong>of</strong> new social-economic <strong>an</strong>d<br />

industrial facilities, residential houses <strong>an</strong>d related infrastructure. It is also largely covered by the<br />

potentially exp<strong>an</strong>sive <strong>an</strong>d reactive, relatively unstable <strong>an</strong>d problematic black cotton soils. Previous<br />

works <strong>an</strong>d studies in adjacent areas <strong>an</strong>d sections <strong>of</strong> the present area (Anyumba, 1991; M<strong>an</strong>ga,<br />

1988; Mowlem, 1985), together with recent field observations by the author reveal problems <strong>of</strong><br />

ground instabilities, failures <strong>an</strong>d movements in areas <strong>of</strong> black cotton soils with serious <strong>engineering</strong>,<br />

environmental <strong>an</strong>d social-economic implications.<br />

The field study phase <strong>of</strong> this project was carried out in Nairobi (Kenya) in the period <strong>of</strong> April-<br />

September, 1999; <strong>an</strong>d December, 2000-J<strong>an</strong>uary, 2001. The laboratory study phase was undertaken<br />

in the period <strong>of</strong> November, 1999-March, 2001; based at the Department <strong>of</strong> Engineering Geology,<br />

Institute <strong>of</strong> Geology <strong>an</strong>d Palaeontology, Technical University <strong>of</strong> Clausthal-Germ<strong>an</strong>y. The field<br />

methods <strong>an</strong>d laboratory procedures followed in the investigation, study <strong>an</strong>d <strong>an</strong>alysis <strong>of</strong> the soils <strong>an</strong>d<br />

underlying bedrock are summarised in the relev<strong>an</strong>t sections in later chapters.<br />

Adequate <strong>an</strong>d proper <strong>engineering</strong> <strong>geological</strong> studies <strong>an</strong>d investigations should serve to provide<br />

useful <strong>an</strong>d detailed information on aspects <strong>of</strong> geology (rock types <strong>an</strong>d lithostratigraphy), types <strong>of</strong><br />

soil, geomorphology <strong>an</strong>d <strong>geological</strong> structures, hydrology <strong>an</strong>d hydrogeology, <strong>an</strong>d geotechnical<br />

characteristics <strong>an</strong>d potential behaviour <strong>of</strong> rocks <strong>an</strong>d/ or soils; followed by <strong>an</strong> assessment <strong>of</strong> potential<br />

instabilities <strong>an</strong>d possible natural hazards, <strong>an</strong>d a subsequent consideration <strong>of</strong> suitable development<br />

projects (Attewell, 1976; Bell <strong>an</strong>d Pettinga, 1985; Blyth, 1974; Hedberg 1976; Sowers, 1994;<br />

Tilford, 1994). In addition to the above investigations, this study also includes a detailed chemical<br />

<strong>an</strong>d mineralogical <strong>an</strong>alysis <strong>an</strong>d evaluation <strong>of</strong> the soils <strong>of</strong> the area. Genetic aspects <strong>of</strong> the soils are<br />

also discussed.


2<br />

In most developing countries including Kenya, there are difficulties <strong>of</strong> undertaking <strong>an</strong>d / or<br />

executing <strong>an</strong> exhaustive <strong>an</strong>d comprehensive soils investigation programme time <strong>an</strong>d again, due to<br />

economic constraints related to limited fin<strong>an</strong>ce as well as lack <strong>of</strong> adequate expertise <strong>an</strong>d/ or<br />

equipment. In such cases, the use <strong>of</strong> soil index properties to infer for soil reactivity <strong>an</strong>d stability<br />

characteristics could be partly a solution to the problem. This study attempts to correlate soil index<br />

parameters with those related to shrink-swell potential, swelling pressures, shear strength <strong>an</strong>d<br />

consolidation-settlement (compressibility); in <strong>an</strong> effort to find closely approximating <strong>an</strong>d/ or fitting,<br />

<strong>an</strong>d therefore predictive relationships between them.<br />

Engineering <strong>geological</strong> investigations <strong>an</strong>d mapping <strong>of</strong> soils carried out in this study was in<br />

accord<strong>an</strong>ce with known st<strong>an</strong>dard practices <strong>an</strong>d/ or procedures (IAEG, 1981a/ b; IAEG, 1976 ).<br />

1.2 Location <strong>of</strong> the study area<br />

Nairobi city is currently bounded by latitudes 1°15´S <strong>an</strong>d 1°20´S to the north <strong>an</strong>d south ; <strong>an</strong>d by<br />

longitudes 36°45´E <strong>an</strong>d 36°55´E to the west <strong>an</strong>d east; respectively (Fig. 1.1). It is the largest town<br />

in Kenya <strong>an</strong>d is located about 30 km to the east <strong>of</strong> the north - south running Great Rift Valley<br />

(Saggerson, 1991; <strong>an</strong>d Fig. 3.1). It extends for about 18,7 km east - west <strong>an</strong>d about 9,3 km north -<br />

south, covering <strong>an</strong> area <strong>of</strong> about 174 km² (Nairobi City Council, 1999). The western <strong>an</strong>d northwestern<br />

sections <strong>of</strong> the city are a part <strong>of</strong> the Kikuyu highl<strong>an</strong>ds that rise from <strong>an</strong> altitude <strong>of</strong> about<br />

1660 m at the City centre to over 1800 m above sea level on the immediate eastern fl<strong>an</strong>ks <strong>of</strong> the Rift<br />

Valley. These sections are covered by thick deposits <strong>of</strong> red soils that increase in their depth<br />

westwards <strong>an</strong>d north-westwards from about 4 m at the City centre to over 10 m in the vicinity <strong>of</strong> the<br />

Rift Valley, outside the present area. The soils are underlain by a series <strong>of</strong> Tertiary volc<strong>an</strong>ic rocks, a<br />

result <strong>of</strong> several episodes <strong>of</strong> volc<strong>an</strong>ic activity that accomp<strong>an</strong>ied the formation <strong>of</strong> the Great Rift<br />

Valley (Fairburn, 1963; Gregory, 1921; Saggerson, 1991; Sikes, 1939).<br />

The southern <strong>an</strong>d eastern sections <strong>of</strong> the city are a part <strong>of</strong> the generally flat to gently rolling Athi<br />

<strong>an</strong>d Kapiti plains. The plains are covered by thick surficial deposits <strong>of</strong> black cotton soils that are<br />

underlain mainly by volc<strong>an</strong>ic rocks; while isolated exposures <strong>of</strong> Basement system Precambri<strong>an</strong><br />

metamorphic rocks also occur (Saggerson, 1991).<br />

The present study area is largely located to the immediate south <strong>of</strong> Nairobi metropolit<strong>an</strong> which is a<br />

part <strong>of</strong> the Athi <strong>an</strong>d Kapiti plains. The remaining part <strong>of</strong> the area to the north forms the southern<br />

half <strong>of</strong> Nairobi city. The whole project area is bounded by latitudes 1°15´S <strong>an</strong>d 1°24´S to the north<br />

<strong>an</strong>d south; <strong>an</strong>d longitudes 36°45´E <strong>an</strong>d 36°58´E to the west <strong>an</strong>d east ; respectively. It occupies <strong>an</strong><br />

area <strong>of</strong> about 367,38 km² in a zone <strong>of</strong> mainly black cotton soils; <strong>an</strong>d to a lesser extent, red soils<br />

which are confined in the north-western section <strong>of</strong> the area (Scott, 1963; Surv. <strong>of</strong> Kenya, 1990).<br />

1.3 Definition <strong>of</strong> study problem<br />

The undertaking <strong>of</strong> the present research study has been prompted by a number <strong>of</strong> factors, <strong>an</strong>d<br />

include:-<br />

- ch<strong>an</strong>ging residential construction practice<br />

- widespread damage to roads/ pavements, residential <strong>an</strong>d low rise buildings as a result <strong>of</strong><br />

exp<strong>an</strong>sive movements in areas <strong>of</strong> black <strong>clays</strong><br />

- lack <strong>of</strong> sufficient data to facilitate <strong>an</strong> <strong>engineering</strong> <strong>characterisation</strong> <strong>of</strong> the soils found<br />

- predicted urb<strong>an</strong> population growth <strong>of</strong> the Nairobi city<br />

- predicted growth <strong>of</strong> services <strong>an</strong>d industrial activities in the city<br />

- predicted growth <strong>an</strong>d exp<strong>an</strong>sion <strong>of</strong> the Nairobi city into its surrounding areas


Figure1.1 Map <strong>of</strong> Nairobi region showing location <strong>of</strong> present study area (Ot<strong>an</strong>do, 2003; Surv. <strong>of</strong> Kenya, 1991).<br />

3


4<br />

1.4 Summary <strong>of</strong> case problems<br />

The Nairobi city has a current population <strong>of</strong> about 3 million people (Nairobi City Council, 1999). It<br />

is also characterised by a number <strong>of</strong> social - economic activities which involve construction <strong>of</strong><br />

residential housing <strong>an</strong>d industrial facilities. Other activities involve construction <strong>of</strong> roads, rail lines,<br />

bridges, airports <strong>an</strong>d sports facilities; power <strong>an</strong>d telephone lines; water <strong>an</strong>d oil pipelines; as well as<br />

sewerage lines. Tunnelling <strong>an</strong>d dam projects are also found. The population <strong>of</strong> the city is expected<br />

to grow in the coming years. It is also expected that the above - mentioned activities will exp<strong>an</strong>d<br />

<strong>an</strong>d/ or multiply, while totally new ones may also be introduced. It is therefore <strong>an</strong>ticipated that there<br />

will be increasing pressure on currently available space <strong>an</strong>d facilities in the city, prompting the need<br />

by the Nairobi City Council <strong>an</strong>d City pl<strong>an</strong>ners to consider exp<strong>an</strong>ding into adjacent areas.<br />

In the past, residential construction on the outskirts <strong>of</strong> Nairobi city, which include the present study<br />

area, were predominated by structures constructed out <strong>of</strong> timber <strong>an</strong>d iron sheets, with strip/ pad<br />

footings. However, the recent ten years have witnessed strict regulations put in place by the Nairobi<br />

City Council aimed at encouraging construction <strong>of</strong> more durable low-rise buildings <strong>an</strong>d bungalows<br />

(Plates 1.1 (a) & (b)) on slab foundations using dimension stones <strong>an</strong>d / or bricks to cater for the<br />

increasing urb<strong>an</strong> population. This shifting trend towards masonry construction justifies a close<br />

study, examination, <strong>engineering</strong> properties <strong>characterisation</strong> <strong>an</strong>d assessment <strong>of</strong> the potential<br />

capabilities <strong>an</strong>d/ or limitations <strong>of</strong> the soils found, especially the exp<strong>an</strong>sive black cotton soils.<br />

Plates 1.1 (a) & (b) New stone structures coming up in the study area.<br />

The current project area is therefore also a potential area <strong>of</strong> exp<strong>an</strong>sion <strong>of</strong> the Nairobi metropolit<strong>an</strong><br />

due to its proximity to the city. It is especially favoured by a number <strong>of</strong> major communication lines<br />

already traversing it <strong>an</strong>d include the north - south running Nairobi - Mombasa road <strong>an</strong>d the Kenya -<br />

Ug<strong>an</strong>da railway; while the Jomo Kenyatta International Airport <strong>an</strong>d Wilson Aiport are also found<br />

within. However, the ambitious pl<strong>an</strong>s by the Nairobi City Council to locate new <strong>an</strong>d light<br />

constructed facilities in the study area may be complicated by the fact that the area is covered by<br />

black cotton soils <strong>an</strong>d also receives heavy rainfall during the wet season. These soils are highly<br />

exp<strong>an</strong>sive <strong>an</strong>d reactive, <strong>an</strong>d would usually exhibit very strong shrink-swell capabilities during<br />

alternating dry <strong>an</strong>d wet months <strong>of</strong> the year, with accomp<strong>an</strong>ying general lowering or rise <strong>of</strong> the<br />

ground surface, respectively (Mitchell,1993; Nelson <strong>an</strong>d Miller, 1992).


5<br />

Plates 1.2 (a) & (b) Strong shrinkage cracks in black cotton soils (black <strong>clays</strong>) during dry months.<br />

Examples <strong>of</strong> structural damage triggered <strong>of</strong>f by ground movements associated with exp<strong>an</strong>sive <strong>an</strong>d<br />

reactive soils in the Nairobi area include cracked pavements as well as cracked <strong>an</strong>d peeled-<strong>of</strong>f<br />

asphalt on tarmac roads giving way to formation <strong>of</strong> pot holes (Plates 1.3 (a) & (b)). The cause is<br />

strong shrinkage cracks (Plates 1.2 (a) & (b)) that usually characterise black cotton soils during the<br />

dry months. In light residential <strong>an</strong>d low-rise buildings, the instabilities m<strong>an</strong>ifest themselves in the<br />

form <strong>of</strong> cracked floors <strong>an</strong>d walls <strong>an</strong>d/ or tilting buildings (Plate 1.4). As part <strong>of</strong> <strong>an</strong> effort to <strong>of</strong>fset<br />

the huge costs the government has to incur every year in rehabilitating <strong>an</strong>d maintaining damaged<br />

roads, pavements <strong>an</strong>d other infrastructure, the ministries <strong>of</strong> Environment <strong>an</strong>d Public Works are<br />

emphasising on soil research aimed at providing data <strong>an</strong>d information relev<strong>an</strong>t to producing suitable<br />

foundation designs capable <strong>of</strong> withst<strong>an</strong>ding the otherwise destructive effects <strong>of</strong> the reactive black<br />

cotton soils found. Best suited <strong>an</strong>d economical ground <strong>an</strong>d/ or soil stabilisation techniques are also<br />

to be sought. This study therefore forms a part <strong>of</strong> the collective efforts geared towards fulfilling<br />

these goals.<br />

Plate 1.3 (a) Peeling asphalt on tarmac road.<br />

Plate 1.3 (b) Pothole formed on tarmac road.<br />

Light <strong>an</strong>d surface <strong>engineering</strong> structures are usually constructed on/ or within the upper layers <strong>of</strong><br />

soils (Galster, 1977; Maharaj, 1995; Plate 1.5). Because <strong>of</strong> the above stated potential instability<br />

problems that would be posed to such structures in the current study area, a prior systematic <strong>an</strong>d<br />

sound mapping, <strong>characterisation</strong> <strong>an</strong>d evaluation <strong>of</strong> the black cotton soils found together with the<br />

underlying bedrock, just as was undertaken in this study, are <strong>an</strong> essential prerequisite. The results <strong>of</strong><br />

study <strong>an</strong>d investigations so performed would be crucial to a sound <strong>engineering</strong> <strong>geological</strong><br />

assessment <strong>of</strong> the soils. These results would assist geo-engineers <strong>an</strong>d city pl<strong>an</strong>ners to <strong>an</strong>ticipate in<br />

adv<strong>an</strong>ce potential limitations <strong>an</strong>d <strong>engineering</strong> behaviour <strong>of</strong> the soils at sites intended for selected<br />

projects. The possible alternatives at this stage would be for the geo-engineers <strong>an</strong>d pl<strong>an</strong>ners to use<br />

the ground information <strong>an</strong>d data so provided to recommend appropriate ground stabilisation/<br />

improvement techniques aimed at improving the capacity <strong>of</strong> the soils to support increased loads


6<br />

from intended structures; <strong>an</strong>d/ or design <strong>an</strong>d construct suitable foundation structures capable <strong>of</strong><br />

withst<strong>an</strong>ding destabilising effects that may be caused by exp<strong>an</strong>sive ground movements.<br />

Plate 1.4 Slightly tilting structure (towards the background) in a zone <strong>of</strong> black <strong>clays</strong>.<br />

The available <strong>geological</strong> <strong>an</strong>d soils maps covering the study area, are at their best, medium scale. For<br />

<strong>engineering</strong> <strong>geological</strong> purposes, they c<strong>an</strong> be described as regional <strong>an</strong>d therefore less useful <strong>an</strong>d<br />

inadequate. Apart from presenting information mostly on lithology, <strong>geological</strong> structure <strong>an</strong>d<br />

stratigraphy <strong>of</strong> rock formations found, the maps show <strong>an</strong>d give little on surficial deposits <strong>an</strong>d soils.<br />

There is also <strong>an</strong> obvious <strong>an</strong>d total omission <strong>of</strong> information on geotechnical characteristics <strong>of</strong> the<br />

soils <strong>an</strong>d rocks. Within the current study, small scale <strong>engineering</strong> <strong>geological</strong> maps that are site<br />

specific have been developed to provide <strong>an</strong>d present import<strong>an</strong>t ground information <strong>an</strong>d data on soils<br />

in a continouos spatial form interpolated <strong>an</strong>d extrapolated over the project area. The small - scale<br />

detailed <strong>engineering</strong> <strong>geological</strong> soil maps will contain information sorted out by geotechnical<br />

criteria for easy interpretation <strong>an</strong>d readily application by geo - engineers <strong>an</strong>d pl<strong>an</strong>ners. The use <strong>of</strong><br />

these maps <strong>an</strong>d application <strong>of</strong> contained information could also be extended to surrounding areas<br />

with similar lithology.<br />

Plates 1.5 (a) & (b) Foundation structures <strong>of</strong> light <strong>an</strong>d low rise buildings located in black <strong>clays</strong> with<br />

impeded drainage (foreground <strong>of</strong> Plate(b) with stagn<strong>an</strong>t water).


7<br />

Available literature in the form <strong>of</strong> <strong>geological</strong> reports (Fairburn, 1963; Gregory, 1921; Saggerson,<br />

1991; Sikes, 1939) is also regional in approach <strong>an</strong>d dwell mainly on rock formations, while the<br />

soils are merely described as recent to present surficial deposits. Two reports are available on soils<br />

(Gethin - Jones, 1952; Scott, 1963) but the aspects contained are discussed <strong>an</strong>d treated from <strong>an</strong><br />

agricultural perspective <strong>an</strong>d therefore almost entirely irrelev<strong>an</strong>t <strong>an</strong>d <strong>of</strong> little value to the needs <strong>of</strong><br />

geo-engineers <strong>an</strong>d city pl<strong>an</strong>ners. Site investigation reports from isolated areas over the city are also<br />

available (Anyumba, 1991; M<strong>an</strong>ga, 1988; Mowlem, 1985). They report mainly on the state <strong>an</strong>d<br />

condition <strong>of</strong> the underlying bedrock, the investigations having originally been me<strong>an</strong>t for more<br />

versatile <strong>engineering</strong> projects involving heavy foundation loads (Plates 1.6 (a) & (b)). As a result,<br />

little useful information, if <strong>an</strong>y, on geotechnical characteristics <strong>of</strong> soils is included.<br />

Plates 1.6 (a) & (b) Heavy structures/ buildings in a zone <strong>of</strong> black <strong>clays</strong> in the study area.<br />

The present study <strong>an</strong>d research therefore aimed to investigate, assess <strong>an</strong>d establish the <strong>engineering</strong><br />

<strong>geological</strong> characteristics <strong>of</strong> soils, <strong>an</strong>d especially as regards the potentially exp<strong>an</strong>sive <strong>an</strong>d reactive<br />

black cotton soils in the present area, for geotechnical purposes. Attempts were made to present the<br />

results <strong>of</strong> the investigations in geotechnical criteria <strong>an</strong>d/ or such a similar form that would be readily<br />

interpreted <strong>an</strong>d applied by geo-engineers <strong>an</strong>d city pl<strong>an</strong>ners.<br />

Best fitting <strong>an</strong>d/ or closely approximating models to describe, present <strong>an</strong>d predict distribution <strong>an</strong>d<br />

spatial variation <strong>of</strong> import<strong>an</strong>t ground information <strong>an</strong>d <strong>engineering</strong> soil properties across the study<br />

area have been developed. Use was made <strong>of</strong> systematically determined soil parameters, both<br />

laterally as well as with depth, across the area.<br />

Indirect <strong>characterisation</strong> <strong>of</strong> <strong>engineering</strong> soil behaviour in situ (shear strength, compressibility,<br />

shrink-swell potential, swelling pressures) has also been made through correlation between results<br />

<strong>of</strong> index tests performed on disturbed <strong>an</strong>d part samples on one h<strong>an</strong>d, <strong>an</strong>d those <strong>of</strong> the above<br />

mentioned <strong>engineering</strong> properties performed on core <strong>an</strong>d/ or undisturbed samples, on the other<br />

h<strong>an</strong>d. The modelled results may be usefully applied by geo-engineers <strong>an</strong>d city pl<strong>an</strong>ners in readily<br />

<strong>an</strong>ticipating probable geotechnical soil behaviour with respect to projected structures at chosen<br />

sites, based on results <strong>of</strong> soil index tests. Possible forms <strong>of</strong> interrelationship between soil index<br />

properties <strong>an</strong>d Atterberg limits were also investigated.<br />

In short therefore, this study embarked on accumulating relev<strong>an</strong>t site specific information across the<br />

study area. The information involved soil properties as well as environmental characteristics such as<br />

climate, groundwater conditions <strong>an</strong>d vegetation. The approach followed in site <strong>characterisation</strong><br />

included among others the examination, assessment <strong>an</strong>d interpretation <strong>of</strong> the perform<strong>an</strong>ce <strong>of</strong><br />

existing constructed structures <strong>an</strong>d buildings; examination, description <strong>an</strong>d classification <strong>of</strong> soil<br />

pr<strong>of</strong>iles; as well as computation <strong>an</strong>d prediction <strong>of</strong> potential ground movements


8<br />

<strong>an</strong>d soil behaviour, based on results <strong>of</strong> field <strong>an</strong>d laboratory investigations. The investigations would<br />

be specific in seeking to establish <strong>an</strong>d/ or model moisture variation with depth, as well as providing<br />

details on depth <strong>of</strong> cracking, soil reactivity in terms <strong>of</strong> shrink-swell potential, active depth; <strong>an</strong>d<br />

other essential variables necessary in the assessment <strong>an</strong>d/ or computation <strong>of</strong> potential ground<br />

surface movement.<br />

In conclusion, this study sought to investigate <strong>an</strong>d present a background on the nature <strong>of</strong> exp<strong>an</strong>sive<br />

<strong>an</strong>d/ or reactive soils <strong>of</strong> the study area, by following suitable field procedures <strong>an</strong>d sampling criteria<br />

as well as laboratory chemical/ mineralogical <strong>an</strong>d <strong>engineering</strong> soils testing methods. It also sought<br />

to present results <strong>of</strong> soil classifications <strong>an</strong>d correlations between soil index <strong>an</strong>d/ or <strong>engineering</strong><br />

properties. Results <strong>of</strong> statistical <strong>an</strong>alyses involving hypothesis testing seeking to determine the<br />

degree <strong>of</strong> homogeneity or uniformity <strong>of</strong> the soils found, have also been presented. The undertaking<br />

<strong>of</strong> this project by the author was also <strong>an</strong> initiative aimed at providing <strong>an</strong> approach <strong>of</strong> study that<br />

would serve as a basis <strong>an</strong>d/ or framework for comprehensive <strong>an</strong>d detailed <strong>engineering</strong> <strong>geological</strong><br />

investigation <strong>an</strong>d assessment <strong>of</strong> exp<strong>an</strong>sive <strong>an</strong>d reactive soils in lithologically similar areas around<br />

Nairobi, <strong>an</strong>d Kenya at large.<br />

1.5 Physiography<br />

The <strong>geological</strong> history <strong>of</strong> Nairobi, including the current study area, has been dominated by volc<strong>an</strong>ic<br />

activity since Miocene times. These areas are currently underlain by a series <strong>of</strong> volc<strong>an</strong>ic rocks as a<br />

result <strong>of</strong> successive lava flows that originated from centres <strong>an</strong>d fissures on the high eastern fl<strong>an</strong>k <strong>of</strong><br />

the Rift region to the west (Gregory, 1921; Saggerson, 1991; Fig. 3.1; Plates 1.7 (a) & (b)).<br />

Plates 1.7 (a) & (b) View <strong>of</strong> the Great Rift Valley from its eastern fl<strong>an</strong>k, to the west <strong>of</strong> Nairobi area.<br />

The flow <strong>of</strong> lava was generally eastwards, onto a warped <strong>an</strong>d partly dissected pre-Miocene erosion<br />

surface below which were older crystalline metamorphic rocks. A few isolated occurrences <strong>of</strong> rocks<br />

<strong>of</strong> Precambri<strong>an</strong> <strong>an</strong>d Quaternary age are also found.<br />

The geomorphological evolution <strong>of</strong> Nairobi <strong>an</strong>d the study area has therefore been mainly controlled<br />

by the volc<strong>an</strong>ic activity that accomp<strong>an</strong>ied the Rift Valley formation. The physiography <strong>of</strong> these<br />

areas is consequent upon the volc<strong>an</strong>ic rocks found, <strong>an</strong>d the tectonic movements which have affected<br />

them (Saggerson, 1991; Sikes, 1926).<br />

The study area is a part <strong>of</strong> the lava plains (Plate 1.8 (a) & (b))that are bordered to the north-west <strong>an</strong>d<br />

west by the Kikuyu highl<strong>an</strong>ds, <strong>an</strong> extension <strong>of</strong> the high ground <strong>of</strong> the eastern fl<strong>an</strong>k <strong>of</strong> the Rift<br />

Valley; <strong>an</strong>d to the south-west by the Ngong hills. The Kikuyu highl<strong>an</strong>ds are characterised by a steep<br />

downward slope in <strong>an</strong> easterly direction. The drainage is consequent, follows the slope <strong>an</strong>d<br />

comprises <strong>of</strong> a number <strong>of</strong> streams giving rise to parallel ridges, some <strong>of</strong> which get broader as the<br />

streams converge. The streams are still down-cutting giving rise to ridges with convex to uniform


9<br />

slopes. The northern boundary between the two physiographic units <strong>of</strong> the Kikuyu highl<strong>an</strong>ds <strong>an</strong>d<br />

the plains is roughly along the east - west line across the Nairobi city centre (Sikes, 1926). The<br />

plains are made up <strong>of</strong> two parts, i.e. the Athi plains <strong>an</strong>d the northern section <strong>of</strong> the Kapiti plains,<br />

both <strong>of</strong> which extend further southwards <strong>an</strong>d eastwards, respectively, beyond the study area. These<br />

plains are underlain by a succession <strong>of</strong> lava flows which usually alternate <strong>an</strong>d/ or intermingle with<br />

lake beds, stream deposits, tuffs <strong>an</strong>d volc<strong>an</strong>ic ash (Saggerson, 1991).<br />

Plate 1.8 (a) Flat plains around Embakasi station. Plate 1.8 (b) Flat plains around J.K.I.A. Airport.<br />

The Athi plains rise from <strong>an</strong> altitude <strong>of</strong> about 1500 m at Athi River, located about 8 km south-east<br />

<strong>of</strong> the present area; to about 1800 m above sea level farther west in the faulted region near Ngong.<br />

The plains form gently rolling grassl<strong>an</strong>ds with a fairly even surface. They end in <strong>an</strong> abrupt<br />

escarpment just west <strong>of</strong> the Athi river. The abrupt escarpment represents the end <strong>of</strong> a lava flow. In<br />

addition, some smaller <strong>an</strong>d low east-facing scarps also occur within, <strong>an</strong>d represent the margins <strong>of</strong><br />

partly eroded lava flows (Gregory, 1921; Saggerson, 1991). The plains are being dissected by<br />

down-cutting rivers (over 30 m below the general l<strong>an</strong>d surface in some places) currently flowing<br />

into the main Athi river. Some steep-sided valleys are also found in the form <strong>of</strong> gullies <strong>an</strong>d c<strong>an</strong>yonlike<br />

gorges cutting into lavas such as along the Nairobi, Mathare <strong>an</strong>d Gitathuru river valleys in the<br />

northern part, as well as the tributaries <strong>of</strong> the Athi river which include Mokoyeti, Sosi<strong>an</strong> <strong>an</strong>d<br />

Mbagathi river valleys in the southern part <strong>of</strong> the study area, respectively. On the other h<strong>an</strong>d, areas<br />

with s<strong>of</strong>t material are characterised by relatively wide valleys, <strong>an</strong>d this is noticeable in lower<br />

sections <strong>of</strong> the Nairobi river valley, as well as along other stream courses in the southern <strong>an</strong>d southwestern<br />

parts <strong>of</strong> the study area. The plains are mainly underlain by phonolitic volc<strong>an</strong>ic rocks <strong>an</strong>d<br />

tuff (Saggerson, 1991).<br />

The Kapiti plains occur to the east <strong>of</strong> the Athi river, where the topography is more variable, from<br />

gently undulating to flat with a number <strong>of</strong> small hills protruding above the general l<strong>an</strong>d surface.<br />

One large hill mass, the Lukenya, is found located in a metamorphic region <strong>of</strong> the plains to the east<br />

<strong>of</strong> the present area. Underlying the plains are volc<strong>an</strong>ic rocks, tuff <strong>an</strong>d metamorphic rocks. The<br />

plains are located in a drier region having no perm<strong>an</strong>ent rivers; <strong>an</strong>d the seasonal stream-beds found<br />

contain water for only short periods <strong>of</strong> time in a year. The dissection <strong>of</strong> these plains is therefore<br />

relatively less severe. Gully erosion is, however, common <strong>an</strong>d characteristic.<br />

The Athi <strong>an</strong>d Kapiti plains extend farther south-eastwards <strong>an</strong>d eastwards beyond the present study<br />

area, where they are bounded by the central hill masses <strong>of</strong> Machakos District (Machakos hills).<br />

These consist <strong>of</strong> the Ith<strong>an</strong>ga, Mua <strong>an</strong>d Iveti hills as well as the K<strong>an</strong>zalu r<strong>an</strong>ge <strong>an</strong>d Donyo Sabuk.<br />

The hills rise abruptly above the surrounding plains. They are also steep-sided <strong>an</strong>d being actively<br />

dissected, partly as a result <strong>of</strong> existing poor l<strong>an</strong>d m<strong>an</strong>agement.<br />

There is also a general rise in relief from the south-east towards the north-west, across the study<br />

area. Elevations <strong>of</strong> 1540 - 1600 m above sea level occur in the south-eastern parts; 1620 - 1680 m


10<br />

above sea level from Jomo Kenyatta International Airport to Wilson Airport; <strong>an</strong>d 1740 - 1800 m<br />

above sea level in the north-western parts <strong>of</strong> the study area; respectively (Survey <strong>of</strong> Kenya, 1990).<br />

More th<strong>an</strong> 80% <strong>of</strong> the study area, including the southern half, has therefore <strong>an</strong> altitude <strong>of</strong> between<br />

1540 m <strong>an</strong>d 1740 m above sea level. An average elevation <strong>of</strong> about 1650 m above me<strong>an</strong> sea level is<br />

recorded for the study area; with the lowest elevations (1540 - 1580 m above sea level) in the southeastern<br />

(around Beacon R<strong>an</strong>ch Farm) <strong>an</strong>d north –eastern parts <strong>of</strong> the study area. The highest<br />

elevations (1740 - 1800 m above sea level) occur at the far western <strong>an</strong>d north-western sections <strong>of</strong><br />

the study area.<br />

The main drainage in Nairobi, including the study area, is consequent upon the regional topography<br />

<strong>an</strong>d prevailing slope <strong>of</strong> the volc<strong>an</strong>ic rocks. The streams are generally easterly-flowing, but a few<br />

cases <strong>of</strong> drainage <strong>an</strong>d flows to the south-east <strong>an</strong>d south also occur (Morg<strong>an</strong>, 1967; Survey <strong>of</strong> Kenya,<br />

1990; Fig. 1.1). The north-western <strong>an</strong>d western parts <strong>of</strong> the study area bordering the Kikuyu<br />

highl<strong>an</strong>ds are characterised by streams that show features <strong>of</strong> having been deeply incised <strong>an</strong>d to have<br />

re-excavated their former courses. According to Saggerson (1991), this could be attributed to the<br />

continued uplift <strong>of</strong> the Rift region during the late Tertiary, <strong>an</strong>d probably early Pleistocene; <strong>an</strong>d also<br />

to the continual deposition <strong>of</strong> lavas <strong>an</strong>d tuffs onto <strong>an</strong> easterly - inclined surface. This is especially<br />

observable in the Nairobi river valley <strong>an</strong>d its main tributaries <strong>of</strong> Kerichwa Kubwa <strong>an</strong>d Kerichwa<br />

Dogo, as well as in the Mathare, Gitathuru <strong>an</strong>d Ngong river valleys. Lavas, welded tuffs <strong>an</strong>d ash<br />

flows therefore poured down the valleys causing their inundation <strong>an</strong>d choking with volc<strong>an</strong>ic rock.<br />

Lateral erosion was limited, <strong>an</strong>d the erosional effect <strong>of</strong> the tuff-flows was mainly concentrated in<br />

the stream-beds where ch<strong>an</strong>nelling <strong>an</strong>d corasion were at maximum, causing re-excavation <strong>of</strong> the<br />

stream valleys.<br />

The uplifting <strong>an</strong>d deposition <strong>of</strong> volc<strong>an</strong>ic materials have therefore given rise to streams that are<br />

characterised by young valleys with steep gradients <strong>an</strong>d narrow <strong>an</strong>d/ or sharp V-shapes in the northwestern<br />

<strong>an</strong>d western parts <strong>of</strong> the study area. In addition, the rapid down-cutting, together with the<br />

relatively s<strong>of</strong>t character <strong>of</strong> the younger volc<strong>an</strong>ic rocks have resulted in the streams flowing in<br />

generally parallel courses, with limited inst<strong>an</strong>ces <strong>of</strong> river capture (Gregory, 1921; Morg<strong>an</strong>, 1967).<br />

M<strong>an</strong>y parts <strong>of</strong> the study area receive rainfall amounts <strong>of</strong> around 900 mm per year, with the rainy<br />

seasons concentrated in the periods <strong>of</strong> mid-March to May; <strong>an</strong>d mid-October to mid-December<br />

(Kenya Meteorological Department, 2001). The streams are most active during these months <strong>of</strong> the<br />

year when rainfall is heavy, <strong>an</strong>d headward erosion <strong>of</strong> gullies <strong>an</strong>d tributaries is common. The<br />

tr<strong>an</strong>sported load is mainly a result <strong>of</strong> erosion <strong>of</strong> the thick soil cover, <strong>an</strong>d this frequently gives rise to<br />

flowing streams <strong>of</strong> red mud (Survey <strong>of</strong> Kenya, 1990).<br />

The study area is characterised by two main drainage basins. The Nairobi river <strong>an</strong>d its tributary<br />

valleys dissect <strong>an</strong>d drain the northern <strong>an</strong>d north-western parts; while the tributaries <strong>of</strong> the Athi river<br />

drain the southern <strong>an</strong>d south-western parts <strong>of</strong> the study area, respectively (Fig. 1.1). The main<br />

tributaries <strong>of</strong> the Athi river include the Sosi<strong>an</strong>, Mokoyeti, Donga <strong>an</strong>d Mbagathi valleys; which are<br />

generally medium – wide rivers, with gentle side-slopes <strong>an</strong>d flat bottoms. Aerial photo -<br />

interpretation <strong>an</strong>d topographic surveys show a generally wide drainage pattern over the study area,<br />

with dist<strong>an</strong>ces <strong>of</strong> 1 to 6 km between adjacent ch<strong>an</strong>nels. The drainage <strong>an</strong>d flow direction are<br />

generally consequent upon the local topography <strong>an</strong>d slope direction, with the Nairobi river basin<br />

draining generally eastwards, while the flow in the Athi river basin is mainly to the south <strong>an</strong>d southeast.<br />

A smaller drainage basin occurs mid-way across the study area, <strong>an</strong>d consists <strong>of</strong> the Ngong<br />

river <strong>an</strong>d its tributary <strong>of</strong> Motoine valley, both <strong>of</strong> which flow generally eastwards. However, apart<br />

from the Nairobi river <strong>an</strong>d Athi river <strong>an</strong>d their main tributaries, most <strong>of</strong> the other stream ch<strong>an</strong>nels<br />

are seasonal <strong>an</strong>d dry during the hot months <strong>of</strong> the year (Saggerson, 1991; Survey <strong>of</strong> Kenya, 1990).


11<br />

1.6 Climate<br />

The Nairobi area <strong>an</strong>d its immediate environments, including the current study area, experience a<br />

type <strong>of</strong> climate characterised by temperatures <strong>an</strong>d rainfall patterns that vary according to location,<br />

altitude <strong>an</strong>d season <strong>of</strong> the year (Survey <strong>of</strong> Kenya, 1990). The generally high altitude (about 1700 m<br />

above sea level) results in a climate that is not quite typically <strong>tropical</strong>, despite the closeness <strong>of</strong> these<br />

areas to the Equator. Generally, temperatures are neither uncomfortably high during the days nor<br />

low at nights (Kenya Meteorological Department, 2001). However, exceptional weather conditions<br />

are also experienced during the hot, dry period shortly before the start <strong>of</strong> rains in March, <strong>an</strong>d is<br />

characterised by relatively high temperatures (up to 28 °C at mid-day), low relative humidity (down<br />

to 10%), <strong>an</strong>d occasional dust-storms caused by the moderately strong easterly wind blowing across<br />

the area at this time <strong>of</strong> the year. The weather is also affected by the occurr<strong>an</strong>ce <strong>of</strong> the south-east<br />

winds in the coastal region <strong>of</strong> Kenya during the period <strong>of</strong> June to September. The latter winds<br />

frequently result in the formation <strong>of</strong> a cap <strong>of</strong> cloud over the high ground immediately east <strong>of</strong> the<br />

Great Rift Valley (including Nairobi <strong>an</strong>d the study area). The persist<strong>an</strong>ce <strong>of</strong> the cloud cover causes<br />

low temperatures during the day (up to a maximum <strong>of</strong> 18° C); <strong>an</strong>d also at nights as well as during<br />

early morning hours (as low as 6° to 8° C). Very cold nights may also occur in the months <strong>of</strong><br />

J<strong>an</strong>uary <strong>an</strong>d February when the sky is clear (Kenya Meteorological Department, 2001).<br />

Table 1.1. Me<strong>an</strong> maximum <strong>an</strong>d minimum temperatures for each month <strong>of</strong> the year over a period <strong>of</strong><br />

50 years, for Nairobi city <strong>an</strong>d the present study area (Kenya Meteorological Department, 2001).<br />

Me<strong>an</strong><br />

Me<strong>an</strong><br />

Months<br />

maximum minimum Me<strong>an</strong> r<strong>an</strong>ge<br />

(°C)<br />

(°C)<br />

(°C)<br />

J<strong>an</strong>uary 26,8 13,1 13,7<br />

February 28 13,4 14,6<br />

March 27,4 14,4 13<br />

April 24,6 14,3 10,3<br />

May 24,1 14,2 9,9<br />

June 23,1 12,6 10,5<br />

July 22,3 11,5 10,8<br />

August 22,7 11,8 10,9<br />

September 25,3 12,2 13,1<br />

October 26,2 13,7 12,5<br />

November 23,6 14,4 9,2<br />

December 25,1 13,8 11,6<br />

Year 24,9 13,3 11,6<br />

Nairobi <strong>an</strong>d the current study area are characterised by me<strong>an</strong> minimum temperatures <strong>of</strong> 11,5° to<br />

14,4 °C; <strong>an</strong>d me<strong>an</strong> maximum temperatures r<strong>an</strong>ging from 22,3° to 28 °C. Low temperatures are<br />

registered during the cool, dry <strong>an</strong>d rather cloudy months <strong>of</strong> June to mid - October; with me<strong>an</strong><br />

maximum temperatures <strong>of</strong> 22,3° to 25,3 °C. High temperatures are recorded in the months <strong>of</strong> mid -<br />

December to mid - March, with me<strong>an</strong> maximum temperatures in the r<strong>an</strong>ge <strong>of</strong> 26,8° to 28 °C. The<br />

weather could be generally described as warm to hot, <strong>an</strong>d dry with sunny days during this latter<br />

period <strong>of</strong> the year. The coldest month is July with <strong>an</strong> average monthly temperature <strong>of</strong> 16,9 °C;<br />

while the hottest month is March with <strong>an</strong> average monthly temperature <strong>of</strong> 20,9 °C. The me<strong>an</strong><br />

monthly r<strong>an</strong>ge across the year is from 9,2 °C in November to 14,6 °C in February. A me<strong>an</strong> <strong>an</strong>nual<br />

temperature <strong>of</strong> 19,5 °C is recorded for the study area; with a me<strong>an</strong> <strong>an</strong>nual r<strong>an</strong>ge <strong>of</strong> 11,6 °C. No<br />

large seasonal ch<strong>an</strong>ges in temperature therefore occur. The highest <strong>an</strong>d lowest temperatures


12<br />

recorded for Nairobi <strong>an</strong>d its adjacent areas over a period <strong>of</strong> 25 years are 32,8° C <strong>an</strong>d 3,9° C,<br />

respectively. A summary <strong>of</strong> temperature statistics over the last 50 years for Nairobi <strong>an</strong>d adjacent<br />

areas is presented in Table 1.1 (Kenya Meteorological Department, 2001).<br />

Nairobi <strong>an</strong>d the present study area are also characterised by wet <strong>an</strong>d dry seasons which show<br />

marked differences in the amounts <strong>of</strong> precipitation. The average <strong>an</strong>nual rainfall is about 900 mm;<br />

but the actual amount in <strong>an</strong>y one year may vary from less th<strong>an</strong> 500 mm to more th<strong>an</strong> 1300 mm per<br />

year (Kenya Meteorological Department, 2001). The distribution <strong>of</strong> rainfall is bimodal, with long<br />

rains being registered in the months <strong>of</strong> mid - March to May ( main rainy season, with 137 to 195<br />

mm per month); <strong>an</strong>d short rains from mid - October to mid - December (secondary rainy season,<br />

with 77 to 114 mm per month). The dates when the rainy seasons actually start <strong>an</strong>d end are very<br />

variable, i.e. the beginning <strong>an</strong>d end <strong>of</strong> a wet season are therefore usually not well defined. The two<br />

rainy seasons coincide approximately with the time <strong>of</strong> ch<strong>an</strong>geover <strong>of</strong> the south-west monsoon <strong>an</strong>d<br />

the north-east monsoon currents which affect eastern Africa (including Kenya) in April <strong>an</strong>d<br />

November, respectively. The dry season is experienced from mid - December to mid - March. The<br />

amount <strong>of</strong> rainfall generally tends to increase with increasing altitude, from about 500 mm per year<br />

around Kitengela (1520 m above sea level) located about 8 km to the south <strong>of</strong> the study area; to just<br />

over 900 mm per year around the Nairobi city centre (1660 m above sea level) in the northern part<br />

<strong>of</strong> the study area, respectively. The precipitation is mainly in the form <strong>of</strong> convectional rainfall,<br />

commonly occurring as late afternoon showers (Survey <strong>of</strong> Kenya, 1990; Kenya Meteorological<br />

Department, 2001). A summary <strong>of</strong> average rainfall (mm) for each month <strong>of</strong> the year, based on<br />

records for 50 years is given in Table 1.2.<br />

Average<br />

Rainfall<br />

(mm)<br />

Me<strong>an</strong> relative humidity<br />

values (%)<br />

9:00 am 3:00 pm<br />

Bright sunshine<br />

(average hours<br />

per day)<br />

Months<br />

J<strong>an</strong>uary 48 79 45 8,9<br />

February 48 74 37 9,5<br />

March 115 82 43 8,7<br />

April 195 86 53 7,0<br />

May 137 85 55 5,7<br />

June 42 85 59 5,8<br />

July 15 83 53 4,4<br />

August 21 85 53 4,2<br />

September 24 82 50 5,9<br />

October 52 80 47 7,0<br />

November 114 81 57 7,0<br />

December 77 83 54 7,9<br />

Year 74 78 51 7<br />

Table 1.2. The average rainfall (mm) based on records for 50 years, me<strong>an</strong> relative humidity values<br />

(%) <strong>an</strong>d amount <strong>of</strong> bright sunshine (average hours per day); for each month <strong>of</strong> the year, for Nairobi<br />

city <strong>an</strong>d the present study area (Kenya Meteorological Department, 2001).<br />

Nairobi <strong>an</strong>d its environs are also located within the path <strong>of</strong> predomin<strong>an</strong>tly easterly winds. The<br />

winds are near the ground <strong>an</strong>d blow across the area , generally between north-east <strong>an</strong>d east from<br />

October to April; <strong>an</strong>d between east <strong>an</strong>d south-east from May to September. Strong winds with<br />

speeds <strong>of</strong> 30 to 40 km/h are common during the dry season, just before the start <strong>of</strong> the long rains in<br />

mid - March; <strong>an</strong>d especially from mid-morning to early afternoon. Wind speeds <strong>of</strong> 15 to 25 km/h<br />

are also recorded during other times <strong>of</strong> the year; while nights are commonly characterised by light<br />

winds.<br />

Nairobi city <strong>an</strong>d its adjacent areas are some 400 km away from the sea <strong>an</strong>d do not therefore<br />

experience the rather unpleas<strong>an</strong>t humid heat that usually characterises <strong>tropical</strong> coastal towns. A very


13<br />

marked daily r<strong>an</strong>ge <strong>of</strong> relative humidity is not uncommon, with the air in the early mornings at/ or<br />

very close to saturation. The afternoons are characterised by a relative humidity <strong>of</strong> about 50%,<br />

although this may sometimes be as low as 10% on clear sunny days especially in February <strong>an</strong>d<br />

March (Kenya Meteorological Department, 2001). Table 1.2 shows a summary <strong>of</strong> me<strong>an</strong> relative<br />

humidity values (%) for each month in a year, both in the mornings (9:00 am) <strong>an</strong>d in the afternoons<br />

(3:00 pm).<br />

Nairobi <strong>an</strong>d its surroundings receive a considerable amount <strong>of</strong> sunshine, averaged at about 7 hours<br />

<strong>of</strong> bright sunshine per day, throughout the year. The quality <strong>of</strong> ultra-Violet radiation has also been<br />

shown to be very high (Kenya Meteorological Department, 2001). The early mornings are <strong>of</strong>ten<br />

cloudy. There is therefore 30% more sunshine in the afternoon th<strong>an</strong> in the morning, with westerly<br />

exposures receiving more insolation th<strong>an</strong> those facing east. There is also considerably more<br />

sunshine during the 6 months that the sun is in the southern hemisphere, th<strong>an</strong> when it is in the north.<br />

However, days with no sunshine at all occasionally occur; <strong>an</strong>d this especially during the rainy<br />

season (mid-March to May), <strong>an</strong>d/ or in the months <strong>of</strong> June, July <strong>an</strong>d August. Average number <strong>of</strong><br />

hours per day <strong>of</strong> bright sunshine for each month <strong>of</strong> the year are summarised in Table 1.2 (Kenya<br />

Meteorological Department, 2001).<br />

1.7 Method <strong>of</strong> survey<br />

The criteria used in demarcating the current project area <strong>an</strong>d selection <strong>of</strong> sampling sites were in<br />

such a way as to accomplish the following:<br />

(i)<br />

(ii)<br />

(iii)<br />

(iv)<br />

(v)<br />

Include areas <strong>of</strong> future residential, industrial <strong>an</strong>d civil <strong>engineering</strong> structural<br />

developments<br />

Include areas currently subjected to exp<strong>an</strong>sive ground movements, pronounced soil<br />

reactivity (strong shrinkage cracking <strong>an</strong>d potentially high exp<strong>an</strong>sive/ swelling character)<br />

<strong>an</strong>d with deteriorating <strong>an</strong>d/ or damaged <strong>engineering</strong> structures<br />

Include a r<strong>an</strong>ge <strong>of</strong> terrain, vegetation <strong>an</strong>d site development conditions<br />

Cover fluvial <strong>an</strong>d swampy soil deposits<br />

Ensure site accessibility during the whole project period<br />

In <strong>an</strong>y case, the project area was selected after consulting with relev<strong>an</strong>t government ministries<br />

(Ministry <strong>of</strong> Environment <strong>an</strong>d Natural Resources, Ministry <strong>of</strong> Public works), local councils <strong>an</strong>d<br />

authorities (Nairobi City Council, Kenya Airports Authority), as well as private comp<strong>an</strong>ies <strong>an</strong>d<br />

l<strong>an</strong>downers.<br />

A desk study phase was undertaken before work in the field commenced. This phase involved the<br />

study <strong>an</strong>d interpretation <strong>of</strong> aerial photographs under a stereoscope (Allum, 1966; Colwell, 1983;<br />

Verstappen, 1980), by drawing boundary lines round the various shade tones, vegetation <strong>an</strong>d relief<br />

patterns (steep slopes, valleys, plains, rocky outcrops). These features were later investigated during<br />

the field study phase by following pl<strong>an</strong>ned field pr<strong>of</strong>iles <strong>an</strong>d road traverses. The patterns which<br />

were found to be related to soil had their boundaries marked/ traced on the photos. The information<br />

was then tr<strong>an</strong>sferred from the photos to the 1:25 000/ 50,000 base maps on return from the field.<br />

The aerial photographs covering the present study area <strong>an</strong>d its environs were provided by the<br />

Survey <strong>of</strong> Kenya, <strong>an</strong>d were printed in 1990.<br />

Observations made during the field survey showed that soil boundaries could easily be delineated<br />

on the aerial photos in areas <strong>of</strong> natural vegetation, perennial crops <strong>an</strong>d wetl<strong>an</strong>ds. Varying shade<br />

tones on the photos were found to correspond to different vegetation types <strong>an</strong>d patterns which in<br />

turn served to reflect soil ch<strong>an</strong>ges across the study area in terms <strong>of</strong> differing texture, depths as well


14<br />

as degree <strong>of</strong> drainage <strong>an</strong>d wetness. The black cotton soils could easily be distinguished <strong>an</strong>d<br />

separated from intercalated patches <strong>of</strong> the dark grey mottled <strong>clays</strong> through the presence <strong>of</strong> a<br />

characteristic shrubby vegetation which appears as specks on the photos in areas <strong>of</strong> the former soil<br />

type.<br />

A field investigation pl<strong>an</strong> <strong>an</strong>d/ or map showing sampling sites is presented in a later chapter (Fig.<br />

4.1). A total <strong>of</strong> 336 survey points <strong>an</strong>d 36 sampling sites were covered along pl<strong>an</strong>ned field pr<strong>of</strong>iles<br />

within black <strong>clays</strong> in the study area. Geotechnical logging <strong>of</strong> soil pr<strong>of</strong>iles was effected at<br />

representative soil sections obtained by h<strong>an</strong>d-excavation as well as those exposed at cut-sections <strong>of</strong><br />

stream courses, roads <strong>an</strong>d quarries. The soil sections facilitated correlation <strong>of</strong> soil pr<strong>of</strong>iles across the<br />

study area. They also enabled determination <strong>of</strong> soil thickness as well as depth position <strong>an</strong>d nature <strong>of</strong><br />

underlying bedrock. A st<strong>an</strong>dard description <strong>of</strong> the nature <strong>of</strong> underlying bedrock (in accord<strong>an</strong>ce with<br />

ISRM, 1978a/ b) is included in a later chapter dealing with the geology <strong>of</strong> the study area.<br />

Undisturbed U100 <strong>an</strong>d disturbed soil sampling was done in the excavated pits. Besides excavated<br />

<strong>an</strong>d / or exposed soil sections, the nature <strong>an</strong>d lithological variation <strong>of</strong> soils were also investigated by<br />

employing soil augering techniques. Cone penetration sounding was also used to determine depth<br />

variation <strong>of</strong> soils across the study area. In situ shear strength characteristics <strong>of</strong> soils were<br />

investigated by me<strong>an</strong>s <strong>of</strong> v<strong>an</strong>e shear testing.<br />

The red soils were also investigated in this study (but to a lesser extent th<strong>an</strong> the black <strong>clays</strong>). Their<br />

inclusion in this study was mainly for comparison purposes. A total <strong>of</strong> 6 excavation, soil sampling<br />

<strong>an</strong>d/ or field investigation sites were covered for these soils.


15<br />

Chapter 2<br />

Previous works<br />

2.1 Summary<br />

A Germ<strong>an</strong> pr<strong>of</strong>essor, Suess (1892) summarised the geology <strong>an</strong>d structure <strong>of</strong> the Eastern Rift<br />

Valley, to the west <strong>of</strong> Nairobi city <strong>an</strong>d the present study area. The Eastern Rift Valley runs<br />

across Kenya from Malawi in the south to the Dead Sea in the north.<br />

Gregory traversed the country <strong>of</strong> Kenya in the period <strong>of</strong> 1896 to 1919. He made import<strong>an</strong>t<br />

early contributions by studying <strong>an</strong>d establishing the geology <strong>of</strong> some selected areas. He also<br />

studied <strong>an</strong>d described the major volc<strong>an</strong>ic series underlying Nairobi <strong>an</strong>d the study area; <strong>an</strong>d<br />

which include the Kapiti phonolites, trachytes, agglomerates, the Nairobi building stone <strong>an</strong>d<br />

the basic lavas <strong>of</strong> the Ngong volc<strong>an</strong>ic centre. He named <strong>an</strong>d designated the Kapiti phonolite<br />

as the oldest unit <strong>of</strong> the series. He published his works <strong>an</strong>d findings two years later (Gregory,<br />

1921) as well as a number <strong>of</strong> papers dealing with various aspects <strong>of</strong> the Rift Valley <strong>an</strong>d its<br />

environs.<br />

Prior, <strong>of</strong> the Mineral Department <strong>of</strong> the British Museum, later examined a large number <strong>of</strong> the<br />

rock specimens collected by Gregory from around the country <strong>an</strong>d which included Kapiti<br />

phonolite <strong>an</strong>d phonolitic quartz trachytes from Nairobi <strong>an</strong>d the current study area. In his<br />

published works (Prior, 1903), he recognised various subdivisions <strong>of</strong> the phonolitic series.<br />

Kunzli (1901) gave <strong>an</strong> account <strong>of</strong> the volc<strong>an</strong>ic rocks <strong>of</strong> the Eastern Rift Valley that included a<br />

description <strong>of</strong> the riebeckite trachytes <strong>an</strong>d phonolitic trachytes <strong>of</strong> the Kikuyu highl<strong>an</strong>ds <strong>an</strong>d<br />

north – western parts <strong>of</strong> the present study area.<br />

In 1905, Maufe, then <strong>of</strong> the Geological Survey <strong>of</strong> Great Britain, conducted a <strong>geological</strong><br />

survey along the railway from Mombasa (in the south) to Kisumu (in the west), through<br />

Nairobi. He later published his works <strong>an</strong>d findings (Maufe, 1908) in which he gave a brief<br />

summary <strong>of</strong> the geology between Nairobi <strong>an</strong>d Kijabe (to the north <strong>of</strong> the study area) by<br />

describing the principal rock types found. He also commented on the nature <strong>of</strong> faulting on the<br />

east fl<strong>an</strong>k <strong>of</strong> the Rift Valley.<br />

Earlier works <strong>an</strong>d brief comments on the geology <strong>of</strong> Nairobi <strong>an</strong>d its surroundings were also<br />

done by Walker (1903), Collie, (1912) <strong>an</strong>d Krenkel (1925). Krenkel described the volc<strong>an</strong>ic<br />

succession found by making reference to specific rock types, based on the earlier works <strong>of</strong><br />

Maufe (1908) <strong>an</strong>d Gregory (1921).<br />

Sikes, then Director <strong>of</strong> Public Works Department, Kenya, conducted underground water<br />

surveys in the Nairobi area in the period <strong>of</strong> 1920 to 1925. He also described the grid – fault<br />

structures caused by late faulting in the areas to the west <strong>of</strong> Nairobi city <strong>an</strong>d in the Rift region.<br />

The results <strong>of</strong> his observations <strong>an</strong>d findings were published in subsequent reports (Sikes 1926,<br />

1934 <strong>an</strong>d 1939). Additional hydrological investigations were undertaken in the years that<br />

followed by staff <strong>of</strong> the Ministry <strong>of</strong> Public Works, Nairobi, who examined in detail boreholes<br />

drilled at the time in the rapidly – growing Nairobi metropolit<strong>an</strong> area. The results <strong>of</strong> this work<br />

were incorporated in a technical report by Gevaerts (1964).<br />

Wayl<strong>an</strong>d, formerly Director <strong>of</strong> Geological Survey Ug<strong>an</strong>da, made a survey <strong>of</strong> materials in<br />

Kenya, including Nairobi <strong>an</strong>d the present study area. He investigated the materials for their


16<br />

possible use as building stone <strong>an</strong>d in the m<strong>an</strong>ufacture <strong>of</strong> bricks <strong>an</strong>d tiles, limes <strong>an</strong>d cements.<br />

His findings were published in a report issued by the Imperial Institute (1926).<br />

Bailey Willis (1936) described the Kikuyu scarp to the west <strong>of</strong> Nairobi, the warped peneplain<br />

around Nairobi <strong>an</strong>d the grid – faulting that has affected the volc<strong>an</strong>ic rocks <strong>of</strong> the Rift valley<br />

region <strong>an</strong>d western parts <strong>of</strong> Nairobi area. Saggerson <strong>an</strong>d Baker (1965) also investigated <strong>an</strong>d<br />

described, among others, the Kikuyu scarp <strong>an</strong>d the warped peneplain in the neighbourhood <strong>of</strong><br />

Nairobi area.<br />

Campbell Smith (1931) described <strong>an</strong>d classified the trachytes <strong>an</strong>d phonolites <strong>of</strong> Nairobi <strong>an</strong>d<br />

its surroundings after examining <strong>an</strong>d <strong>an</strong>alysing a number <strong>of</strong> rock specimens collected from the<br />

areas. His findings were published <strong>an</strong>d incorporated in a report.<br />

Bowen visited Kenya in the 1930´s <strong>an</strong>d made <strong>an</strong> excursion into the Rift Valley. He also<br />

collected specimens <strong>of</strong> volc<strong>an</strong>ic rocks from other parts <strong>of</strong> the country, including Nairobi <strong>an</strong>d<br />

the present study area. In 1937, he published the results <strong>of</strong> his studies in a paper in which he<br />

illustrated his experimental work with personal observations on the lavas.<br />

The nature <strong>of</strong> alkali – feldspars in certain Keny<strong>an</strong> rocks, including the Kapiti phonolite was<br />

investigated by Häkli (1960). The specimens <strong>of</strong> Kapiti phonolite had been previously<br />

collected by a Finnish Expedition in 1952 from the Stony Athi River <strong>an</strong>d Nairobi, including<br />

some parts <strong>of</strong> the study area.<br />

Williams (1967) outlined the geology <strong>of</strong> Nairobi <strong>an</strong>d its surroundings, in its regional setting.<br />

Various lavas have also been collected from the Nairobi succession <strong>an</strong>d dated (Baker et al,<br />

1988). Curtiss, Evernden <strong>an</strong>d Miller <strong>of</strong> the University <strong>of</strong> California, USA, carried out K/Ar –<br />

age dating on rock specimens collected from the Nairobi region. The volc<strong>an</strong>ic rocks were<br />

found to be Upper Miocene to Pleistocene in age (Evernden et al, 1964).<br />

Other studies involving Nairobi <strong>an</strong>d covering the present study area include palaeomagnetic<br />

studies by staff <strong>of</strong> the Department <strong>of</strong> Physics, University <strong>of</strong> Nairobi (Mussett et al, 1964-<br />

1965); <strong>an</strong>d the monitoring <strong>an</strong>d detection <strong>of</strong> seismic activities <strong>an</strong>d possible associated earth<br />

movements, by the Seismology Section, Department <strong>of</strong> Geology, University <strong>of</strong> Nairobi<br />

(1990-2002).<br />

A number <strong>of</strong> other <strong>geological</strong> investigations have been carried out in Nairobi <strong>an</strong>d its environs,<br />

including the study area; but they are recorded as unpublished works by the Mines <strong>an</strong>d<br />

Geological Department, Nairobi. Some <strong>of</strong> the unpublished works include those <strong>of</strong> Binge,<br />

Joubert, Pulfrey <strong>an</strong>d Saggerson (Mines <strong>an</strong>d Geological Department, Nairobi, 2001).<br />

Early works regarding detailed survey <strong>an</strong>d study <strong>of</strong> soils <strong>of</strong> the Nairobi area <strong>an</strong>d its environs<br />

were undertaken by Jones in 1952 <strong>an</strong>d published in 1953. His works were later carried<br />

forward to completion by Scott, <strong>of</strong> the then East Afric<strong>an</strong> Agriculture <strong>an</strong>d Forestry Research<br />

Org<strong>an</strong>isation. In his works, he described a variety <strong>of</strong> soils found as having developed under a<br />

wide r<strong>an</strong>ge <strong>of</strong> climatic conditions (Scott, 1963).<br />

Other works in the study <strong>of</strong> soils were undertaken by Stephen, Bellis <strong>an</strong>d Muir (1956), who<br />

investigated the nature <strong>of</strong> the black cotton soils overlying Nairobi phonolite in the Athi<br />

plains, with the purpose <strong>of</strong> establishing the cause for uneven growth <strong>of</strong> crops. They also noted<br />

that soils from Embakasi area, Nairobi, contained some minerals typical <strong>of</strong> metamorphic<br />

rocks.


17<br />

Table 2.1 Stratigraphic correlation <strong>of</strong> rocks in the Nairobi region (Saggerson, 1991).<br />

SOUTH CENTRAL NORTH<br />

Kajiado Nairobi Kijabe<br />

(Matheson, 1966) (Saggerson, 1968) (Thompson, 1964)<br />

Soils, ashes, alluvium,<br />

Soils <strong>an</strong>d ashes<br />

RECENT Soils loess, gypsiferous beds<br />

_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _<br />

Longonot trachytes<br />

Trachytes<br />

U<br />

Lacustrine deposists<br />

xxxxxxxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxxx<br />

Sediments (K<strong>an</strong>dizi Valley)<br />

Pyroclastics<br />

PLEIST- Olorgesailie Upper Trachyte Division <strong>of</strong> xxxxxxxxxxxxxxxxx<br />

OCENE M Lake Beds rhyolites, trachyrhyolites Trachytes, Kijabe<br />

<strong>an</strong>d obsidi<strong>an</strong>s ( with basalts, welded tuffs<br />

interbedded diatomites)<br />

_ _ _ _ _ _ _ _ _ _ __ _ _<br />

xxxxxxxxxxxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxxxxxxxxxxx<br />

Upper Trachyte Division <strong>of</strong>:<br />

Pyroclastics <strong>an</strong>d<br />

Welded tuffs & pyroclastics, Laikipi<strong>an</strong> basalts<br />

(Sattima Series/<br />

L<br />

Laikipi<strong>an</strong> lavas)<br />

Orthophrye trachyte<br />

Orthophrye trachyte, <strong>an</strong>d<br />

Limuru trachytes<br />

Alkali trachyte<br />

xxxxxxxxxxxxxxxxxxxxxxxxxxxxx<br />

Middle Trachyte Division:<br />

Ol Keju Nero basalts<br />

Tigoni & Karura trachyte<br />

xxxxxxxxxxxxxxxxxxxxxxxxx Ruiru Dam trachyte<br />

Ol Esayeti volc<strong>an</strong>ics Kabete trachyte Sattima Series<br />

Kerichwa valley tuffs<br />

Kerichwa Valley tuffs<br />

PLIOCENE <strong>an</strong>d Ol Doinyo Narok Ol Esayeti volc<strong>an</strong>ics <strong>an</strong>d<br />

agglomerate<br />

_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _<br />

Lower Trachyte Division:<br />

Olorgesailie<br />

Ngong volc<strong>an</strong>ics<br />

Volc<strong>an</strong>ic Series Ol Doinyo Narok agglomerate Laikipi<strong>an</strong> lavas<br />

Ololua trachyte<br />

Nairobi / Kiambu trachytes<br />

Mbagathi trachyte<br />

Sediments & tuffs<br />

Nairobi & K<strong>an</strong>dizi phonolites<br />

Mbagathi phonolitic trachytes<br />

x-x-x-x-x-x-x-x-x-x-x-x-x-x-x-x-x-<br />

Upper Athi tuffs<br />

Athi tuffs <strong>an</strong>d Lake Beds<br />

_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _<br />

Lower Trachyte Division:<br />

Simbara basalt<br />

MIOCENE Kapiti phonolite Kapiti phonolite Simbara Series<br />

_________________________ _____________________________ _____________________<br />

PRECAMB- Basement System<br />

Metamorphic rocks <strong>of</strong> the<br />

RIAN Mozambique Belt not exposed<br />

xxx Faulting _ _ _ Erosion surface ___ Major unconformity


18<br />

Nevill (1961), carried out investigations into the suitability <strong>of</strong> soils from Kabete <strong>an</strong>d Sasumua<br />

areas (located to the north <strong>an</strong>d north-east <strong>of</strong> the present area, respectively) for stabilisation<br />

with Portl<strong>an</strong>d cement <strong>an</strong>d hydrated lime.<br />

In 1968, Saggerson reconstructed the stratigraphical sequence <strong>of</strong> formations <strong>of</strong> Nairobi,<br />

including the present study area. In the sequence, loose surface deposists (soils, ashes,<br />

alluvium, loess, gypsiferous beds) are the youngest, having been formed in recent times;<br />

while the Kapiti phonolite, which extruded in the area in Miocene, is the oldest unit <strong>of</strong> the<br />

volc<strong>an</strong>ic succession. These phonolites overly older Precambri<strong>an</strong> metamorphic rocks <strong>of</strong> the<br />

Mozambique Belt. He also correlated the strata sequence found at Nairobi with those <strong>of</strong><br />

neighbouring areas <strong>of</strong> Magadi, Kajiado <strong>an</strong>d Kijabe (Table 2.1). The results <strong>of</strong> his works were<br />

published later by the Mines <strong>an</strong>d Geological Department, Nairobi (Saggerson, 1991).


19<br />

Chapter 3<br />

Geology<br />

3.1 Introduction<br />

Geological records show that much <strong>of</strong> central Kenya, <strong>of</strong> which Nairobi <strong>an</strong>d the present study<br />

area are part, underwent penepl<strong>an</strong>ation on more th<strong>an</strong> one occasion in the period <strong>of</strong> late<br />

Precambri<strong>an</strong> to Tertiary (Saggerson, 1991). Doming, linear warping <strong>an</strong>d erosion <strong>of</strong> the then<br />

sub-Miocene erosion surface was succeeded by the extrusion <strong>of</strong> the first lava flows which<br />

flooded the eastern plains, including Nairobi <strong>an</strong>d the present area, during mid-Miocene time.<br />

This was followed by intermittent volc<strong>an</strong>icity which continued until Recent time. The<br />

volc<strong>an</strong>icity was genetically associated with tectonic movements <strong>an</strong>d the development <strong>of</strong> the<br />

Rift Valley to the west <strong>of</strong> these areas. The western parts <strong>of</strong> Nairobi <strong>an</strong>d sections <strong>of</strong> the present<br />

study area have also been affected by faulting on at least three occasions, resulting in<br />

spectacular topography. The lavas <strong>an</strong>d pyroclastics originated mainly from fissure eruptions<br />

<strong>of</strong> the Rift region. Two central volc<strong>an</strong>oes, Ngong <strong>an</strong>d Ol Esayeiti, were also active during this<br />

period <strong>an</strong>d contributed their share <strong>of</strong> lavas <strong>an</strong>d pyroclastic material to the eastern plains. The<br />

volc<strong>an</strong>oes are located to the south-west <strong>of</strong> Nairobi city.<br />

The geology <strong>of</strong> Nairobi <strong>an</strong>d the present study area is, therefore, characterised mainly by a<br />

succession <strong>of</strong> volc<strong>an</strong>ic rocks <strong>an</strong>d pyroclastics <strong>of</strong> Cainozoic age (Saggerson, 1991). Underlying<br />

the volc<strong>an</strong>ic rocks is a foundation <strong>of</strong> folded crystalline metamorphic rocks (gneisses <strong>an</strong>d<br />

schists) <strong>of</strong> Precambri<strong>an</strong> age which belongs to the Mozambique Belt. The metamorphic rocks<br />

are only occasionally exposed <strong>an</strong>d some <strong>of</strong> their fragments have been noted in agglomerates<br />

previously derived from the former Ngong volc<strong>an</strong>o, to the west <strong>of</strong> the area.<br />

The volc<strong>an</strong>ic rocks <strong>of</strong> Nairobi <strong>an</strong>d its environs, including the study area, are a part <strong>of</strong> a wider<br />

East Afric<strong>an</strong> alkaline suite characterised by a domin<strong>an</strong>ce <strong>of</strong> soda over potash. The rocks are<br />

distinguished into two groups; the first <strong>of</strong> which is a strongly alkaline series represented by<br />

feldspathoid-bearing phonolites, bas<strong>an</strong>ites, tephrites <strong>an</strong>d more basic varieties. The second<br />

group is a mildly alkaline series <strong>an</strong>d includes feldspathoid-free rocks containing soda-rich<br />

amphiboles <strong>an</strong>d pyroxenes. Differentiation <strong>of</strong> members <strong>of</strong> the two series is accomp<strong>an</strong>ied by<br />

<strong>an</strong> increase in silica content, giving rise to trachytes, rhyolites <strong>an</strong>d obsidi<strong>an</strong>s. On the contrary,<br />

however, volc<strong>an</strong>ic rocks associated with the Western Rift Valley (to the west, outside Kenya)<br />

belong to the potash-rich alkaline series (Saggerson, 1991).<br />

The rock formations are covered <strong>an</strong>d overlain by thick deposits <strong>of</strong> soils <strong>an</strong>d gravels <strong>of</strong><br />

Quaternary age. However, areas within the Rift are characterised by loess <strong>an</strong>d lacustrine<br />

deposits, some containing diatomaceous beds; <strong>an</strong>d all <strong>of</strong> which serve to reflect major ch<strong>an</strong>ges<br />

in climatic conditions (Saggerson, 1991).<br />

A number <strong>of</strong> authors have assigned various ages to the volc<strong>an</strong>ic rocks <strong>of</strong> Nairobi region. A<br />

majority <strong>of</strong> the formations are suggested to be Tertiary to Recent. Gregory, on the other h<strong>an</strong>d,<br />

was <strong>of</strong> the view that the earliest lavas <strong>of</strong> the Kapiti<strong>an</strong> series were Cretaceous (Gregory, 1921).<br />

Potassium-argon age-dating <strong>of</strong> the rocks was done by Curtiss, Evernden <strong>an</strong>d Miller, <strong>of</strong> the<br />

University <strong>of</strong> California. Their findings indicate volc<strong>an</strong>icity as having beg<strong>an</strong> in Upper<br />

Miocene, <strong>an</strong>d had almost completely ceased by the end <strong>of</strong> Pleistocene (Evernden et al, 1964).<br />

The dating was performed on whole rock specimens as well as specific minerals<br />

(<strong>an</strong>orthoclase, s<strong>an</strong>idine, feldspar, nepheline), <strong>an</strong>d shows the ages <strong>of</strong> the rocks as varying from<br />

0,8 Ma (Plateau trachyte series, Magadi road) to 13,40 Ma (Kapiti phonolite, Stony Athi


20<br />

River). The dating also involved rock specimens from Nairobi metropolit<strong>an</strong> <strong>an</strong>d present study<br />

areas, <strong>an</strong>d included Nairobi trachyte (3,17-3,45 Ma), Nairobi stone (4,84-5,67 Ma), Nairobi<br />

phonolite (5,2-10,22 Ma) <strong>an</strong>d Kapiti phonolite (12,90-13,40 Ma). There is also the likelihood<br />

that volc<strong>an</strong>ic activity continued into Recent times in areas to the west <strong>of</strong> Nairobi. This is<br />

evidenced by the occurrence <strong>of</strong> weathered tuffs <strong>an</strong>d ash layers within soil pr<strong>of</strong>iles,<br />

representing the eruptive periods <strong>of</strong> the later times (Saggerson, 1991).<br />

A summarised discussion <strong>of</strong> the major rock types underlying the present study area is given in<br />

the following sections.<br />

3.2 Precambri<strong>an</strong> rocks <strong>of</strong> the Mozambique Belt<br />

The crystalline metamorphic rocks <strong>of</strong> Precambri<strong>an</strong> age outcrop in areas where the volc<strong>an</strong>ic<br />

cover has been removed by erosion. They are especially noticeable in the southern <strong>an</strong>d southwestern<br />

parts <strong>of</strong> the study area where they occur as isolated exposures in shallow stream<br />

courses, or as float on the stream edges. They are generally small highly weathered outcrops<br />

consisting mainly <strong>of</strong> layered fine-grained schists <strong>an</strong>d coarse gneisses that have been invaded<br />

in some places by pink quartzo-feldspathic pegmatites. Also found but to a lesser extent are<br />

biotite-garnet-epidote gneisses, hornblende gneisses <strong>an</strong>d quartz-feldspar gneisses.<br />

The structures revealed in the metamorphic rocks generally indicate north-south trending<br />

folds. However, the rocks also show some variations in strike <strong>an</strong>d dip due to generally smallscale<br />

folds. They also exhibit lineations which plunge at approximately 14° - 21° to the southwest;<br />

<strong>an</strong>d which are most probably related to the north-east to south-west fold trends that have<br />

been recognised in neighbouring areas (Saggerson et al, 1960). The north-south <strong>an</strong>d north-east<br />

to south-west fold trends are typical <strong>of</strong> much <strong>of</strong> southern Kenya, including the Nairobi area,<br />

<strong>an</strong>d serve to point to the fact that the underlying Precambri<strong>an</strong> rocks have been subjected to a<br />

minimum <strong>of</strong> two phases <strong>of</strong> deformation (Saggerson, 1991).<br />

In other areas, the metamorphic rocks are unexposed <strong>an</strong>d separated from the overlying<br />

volc<strong>an</strong>ic rocks by a pl<strong>an</strong>e <strong>of</strong> unconformity which represents <strong>an</strong> erosional surface formed in<br />

early Miocene. Evidences <strong>of</strong> the erosional surface occur in the form <strong>of</strong> thin layers <strong>of</strong><br />

weathered metamorphic materials separating the Precambri<strong>an</strong> <strong>an</strong>d volc<strong>an</strong>ic rocks, <strong>an</strong>d<br />

exposed in some <strong>of</strong> the streams in the southern part <strong>of</strong> the study area. However, cases <strong>of</strong> soil<br />

materials (<strong>clays</strong>, s<strong>an</strong>ds, detritus) directly overlying the Precambri<strong>an</strong> rocks also occur, <strong>an</strong>d is<br />

evidenced by a number <strong>of</strong> borehole records in the present area. According to Sikes (1939), the<br />

presence <strong>of</strong> a former topography across the study area is also revealed by borehole records,<br />

which show the crystalline rocks as occurring at different elevations above sea-level.<br />

Borehole records <strong>of</strong> the Nairobi Station (BH 1937c; 37° 00`E, 1° 24`S), for inst<strong>an</strong>ce, show the<br />

erosional surface to have been penetrated at <strong>an</strong> elevation <strong>of</strong> about 838 m above sea-level. In<br />

other places, the crystalline rocks were intercepted at elevations <strong>of</strong> 1519 m above sea-level<br />

(BH 2655) in the south-eastern part <strong>of</strong> the area; 91 m depth below ground surface at Wattle<br />

Blossom Farm, southern part <strong>of</strong> the area; <strong>an</strong>d 1509 m above sea-level in the area around<br />

Nairobi National Park. Records also reveal the crystalline metamorphic rocks to have<br />

undergone cyclic penepl<strong>an</strong>ations in early Tertiary. Evidence <strong>of</strong> earlier sedimentation is,<br />

however, absent. The latest metamorphic events to affect the rocks are dated at 480-530 Ma,<br />

based on the age <strong>of</strong> Phlogopite in crystalline limestone from Turoka, in southern Nairobi area<br />

(Saggerson, 1991).<br />

Early studies on soil samples collected from around Embakasi in the south-eastern part <strong>of</strong> the<br />

present area (Saggerson et al, 1968; Stephen et al, 1956) showed them to contain typical


21<br />

metamorphic minerals sillim<strong>an</strong>ite, ky<strong>an</strong>ite, staurolite <strong>an</strong>d garnet. This serves to suggest that<br />

some <strong>of</strong> the volc<strong>an</strong>ic material <strong>of</strong> the Nairobi region, which on weathering gave rise to Recent<br />

soils, had during their eruption incorporated pieces <strong>of</strong> the crystalline metamorphic rocks.<br />

3.3 Volc<strong>an</strong>ic rocks<br />

The Nairobi region, including the present study area, are covered by a thick succession <strong>of</strong><br />

alkaline volc<strong>an</strong>ic rocks <strong>an</strong>d associated tuffs, derived from the Rift Valley region (Saggerson,<br />

1991).The rocks accumulated in the period <strong>of</strong> mid-Miocene to atleast Upper Pleistocene,<br />

during which time volc<strong>an</strong>ic activity dominated the <strong>geological</strong> history <strong>of</strong> the Nairobi region.<br />

According to Saggerson (1991), the volc<strong>an</strong>ic rocks <strong>of</strong> Nairobi region are estimated to amount<br />

in volume to more th<strong>an</strong> 1042 km³, <strong>an</strong>d cover <strong>an</strong> area <strong>of</strong> about 3000 km². The rocks are a part<br />

<strong>of</strong> a more extensive late Cainozoic (Pliocene) to Recent petrographic province that extends<br />

outside Kenya, from Ethiopia in the north to T<strong>an</strong>z<strong>an</strong>ia in the south.<br />

The earliest volc<strong>an</strong>ic activity beg<strong>an</strong> with the extrusion <strong>of</strong> the Kapiti phonolite (dated at 13<br />

Ma) onto a dissected erosion surface, which had developed across nearly the entire East<br />

Afric<strong>an</strong> region, in pre-Miocene times (Saggerson <strong>an</strong>d Baker, 1965).Weathered pyroclastic<br />

materials are also found associated in the Nairobi area, occurring as interstratifications within<br />

the lava succession <strong>an</strong>d in the deep red soils. They were originally fragmentary deposits, a<br />

result <strong>of</strong> violent volc<strong>an</strong>ic activity from a number <strong>of</strong> centres within the Rift region, during<br />

Pleistocene (Saggerson, 1991). The only identified vents include those <strong>of</strong> the Ngong volc<strong>an</strong>o<br />

<strong>an</strong>d its satellites, Ol Esayeiti <strong>an</strong>d Ngoroi, all <strong>of</strong> which occur to the south – west <strong>of</strong> Nairobi<br />

city. Much <strong>of</strong> the volc<strong>an</strong>ic material must have, therefore, derived from fissure eruptions from<br />

dispersing centres located in the Kikuyu-Limuru area, to the west <strong>of</strong> Nairobi city <strong>an</strong>d the<br />

present study area. Figure 3.1(a) shows the extent <strong>of</strong> early lava flows in the Nairobi region,<br />

including the study area; while Table 2.1 (in previous chapter) gives a corresponding<br />

stratigraphic classification. The approximate ages <strong>of</strong> rock formations immediately underlying<br />

the soils <strong>of</strong> the area are given in Fig. 3.1(b).<br />

There is a general thinning <strong>of</strong> the lava flows in <strong>an</strong> easterly direction across the areas covered.<br />

The eastern limits <strong>of</strong> the flows are characterised by erosional scarps, which form terminal<br />

walls <strong>of</strong> the more resist<strong>an</strong>t lava sheets. Studies have also shown the volc<strong>an</strong>ic material to be<br />

mainly phonolitic <strong>an</strong>d trachytic in character; while the pyroclastics are domin<strong>an</strong>tly trachytic<br />

(Saggerson <strong>an</strong>d Baker,1965).


24<br />

The following section includes a summary <strong>of</strong> the major volc<strong>an</strong>ic rocks underlying the present<br />

study area.<br />

3.3.1 Kapiti phonolite<br />

Borehole records from Nairobi city <strong>an</strong>d the current study area show the Kapiti phonolites as<br />

underlying associated volc<strong>an</strong>ic rocks. They are also observed to overly a sub-volc<strong>an</strong>ic floor <strong>of</strong><br />

erosion surface previously cut into Precambriam metamorphic rocks. These rocks therefore<br />

represent the oldest unit <strong>of</strong> a succession <strong>of</strong> lavas present in the study area (Gregory, 1921).<br />

Exposures <strong>of</strong> Kapiti phonolite are observable in the south-eastern part <strong>of</strong> the study area,<br />

generally in the form <strong>of</strong> small outcrops confined along some <strong>of</strong> the stream courses found; <strong>an</strong>d<br />

in the southern part <strong>of</strong> the area in the tributary valleys <strong>of</strong> the Athi river. Farther south, outside<br />

the present area, small outcrops also occur in the valley <strong>of</strong> the Athi river <strong>an</strong>d in the tributary<br />

valleys <strong>of</strong> Kitengela river; while extensive outcrops are in the valley <strong>of</strong> the Stony Athi River,<br />

which forms the north-western border <strong>of</strong> the Kapiti plains. The rock unit descends with a<br />

slope nearly identical to that <strong>of</strong> the sub-volc<strong>an</strong>ic floor, from 1585 m above sea-level in the<br />

south-eastern part, to1280 m above sea-level in the northern part <strong>of</strong> the study area <strong>an</strong>d the<br />

Nairobi City centre (Saggerson, 1991).<br />

The lava originated from <strong>an</strong> area to the south <strong>of</strong> the present study area <strong>an</strong>d was most probably<br />

derived from vents located on the fl<strong>an</strong>ks <strong>of</strong> the Rift Valley <strong>an</strong>d/ or from the site <strong>of</strong> the former<br />

Olorgasailie central volc<strong>an</strong>o (Baker, 1958). It was a part <strong>of</strong> the first Miocene flood eruptions<br />

<strong>an</strong>d flowed generally northwards onto the eroded surface, which was covered in places by<br />

Tertiary conglomerates <strong>an</strong>d grits (Fairburn, 1963). The Kapiti lava also extends westwards up<br />

to Lion Drift in the Nairobi National Park by thinning against the higher ground. The Mua<br />

hills <strong>an</strong>d the K<strong>an</strong>zalu r<strong>an</strong>ge located to the east <strong>of</strong> the study area acted as a barrier to the<br />

northerly lava flow, <strong>an</strong>d instead, encouraged a north-westerly flow across the area. A further<br />

contribution originated from areas to the north <strong>of</strong> the study area, where the flows had also<br />

given rise to the formation <strong>of</strong> the Yatta plateau phonolite.<br />

Borehole records from Nairobi metropolit<strong>an</strong> area, the study area <strong>an</strong>d around Athi River<br />

township (south <strong>of</strong> the study area) show the presence <strong>of</strong> sedimentary strata in the form <strong>of</strong><br />

<strong>clays</strong> <strong>an</strong>d other sediments occurring intercalated between <strong>an</strong>d separating several lava flows<br />

(Gevaerts, 1964).<br />

Exposures show parts <strong>of</strong> the phonolite to be vesicular in nature, commonly giving rise to<br />

formation <strong>of</strong> small patches <strong>of</strong> amygdales filled with calcite <strong>an</strong>d zeolite (mainly natrolite). The<br />

elongated feldspar phenocrysts ( up to 76 mm long) <strong>an</strong>d vesicles frequently display a<br />

preferred orientation (Gregory, 1921; Saggerson, 1991).<br />

H<strong>an</strong>d specimens <strong>of</strong> Kapiti phonolite are distinct, <strong>an</strong>d show large crystals <strong>of</strong> feldspar <strong>an</strong>d<br />

waxy-looking nephelines occurring as phenocrysts set in a fine-grained dark green to black or<br />

dark bluish-grey groundmass. The presence <strong>of</strong> phenocrysts <strong>of</strong> these two minerals is<br />

diagnostic, <strong>an</strong>d serves to identify <strong>an</strong>d distinguish the Kapiti phonolite from other lavas across<br />

Kenya (Gregory, 1921).<br />

Thin sections <strong>of</strong> phonolite show phenocrysts <strong>of</strong> <strong>an</strong>orthoclase <strong>an</strong>d nepheline set in <strong>an</strong><br />

interstitial groundmass (Häkli, 1960). Serpentinised pseudomorphs <strong>an</strong>d patches <strong>of</strong> chloritic<br />

material found in the groundmass are probably secondary after olivine, while opaque iron ore


25<br />

<strong>an</strong>d prismatic apatite occur as accessories (Saggerson, 1991). Composition <strong>of</strong> the groundmass<br />

is also summarised in Table (3.4).<br />

Table 3.1 Chemical <strong>an</strong>alyses <strong>of</strong> alkali-feldspars from the Kapiti phonolite <strong>an</strong>d similar rocks.<br />

1 2 3 4<br />

SiO2 65,61 62,79 64,33 64,80<br />

Al2O3 20,48 22,12 20,94 21,14<br />

Fe2O3 0,20 0,36 0,2 0,68<br />

FeO 0,20 0,41 0,58 -<br />

MgO 0,00 - - 0,10<br />

CaO 7,03 3,76 2,01 1,18<br />

Na2O 6,26 7,35 7,22 6,94<br />

K2O 0,32 2,98 4,71 5,26<br />

H2O+ 0,03 0,19 0,27 0,20<br />

H2O- 0,00 0,07 0,10 0,20<br />

Totals 99,93 100,03 100,36 100,00<br />

1. Alkali-feldspar phenocryst (<strong>an</strong>orthoclase). From Kapiti phonolite, Stony Athi River (Häkli, 1960).<br />

2. Loose crystals <strong>of</strong> <strong>an</strong>orthoclase (potash-oligoclase). Crater <strong>of</strong> Mount Erebus, Antarctica (Mountain, 1925).<br />

3. Anorthoclase. Crater slopes, Mount Kenya (Mountain, 1925).<br />

4. Anorthoclase phenocrysts. From Kenyte, Mount Kenya (Spencer, 1937).<br />

The thin section descriptions <strong>an</strong>d results <strong>of</strong> chemical <strong>an</strong>alysis <strong>of</strong> Kapiti phonolite from the<br />

study area <strong>an</strong>d its environs, has been shown by various authors to be similar (Baker, 1954;<br />

Bowen, 1937; Smith Campbell,1950; Saggerson, 1991; Prior, 1903). The results <strong>of</strong> chemical<br />

<strong>an</strong>alyses are summarised in Table 3.1, which also shows a comparison <strong>of</strong> the composition <strong>of</strong><br />

alkali-feldspar phenocrysts from the Kapiti phonolite; with those <strong>of</strong> <strong>an</strong>orthoclase crystals from<br />

Mount Kenya <strong>an</strong>d Antarctica. From the table, specimens from the Kapiti phonolite are found<br />

to show a marked rise in the ratio <strong>of</strong> alkalis, Na2O/K2O, with increase in lime (CaO) content.<br />

Specimens from Mount Kenya <strong>an</strong>d Antarctica, however, exhibit a steady increase in CaO with<br />

increasing Na2O/K2O ratio.<br />

3.3.2 Nairobi phonolite<br />

The Nairobi phonolite covers a large part <strong>of</strong> the Athi plains. It extends from Nairobi National<br />

Park located in the western <strong>an</strong>d south-western parts <strong>of</strong> the study area, <strong>an</strong>d northwards across<br />

the Nairobi city centre to Kiambu township, located north <strong>of</strong> /<strong>an</strong>d outside the present study<br />

area. It overlies the Athi Series with a probable disconformity. Outcrops <strong>of</strong> the phonolite are<br />

observable in the southern <strong>an</strong>d south-western parts <strong>of</strong> the study area in a prominent scarp,<br />

which also shows the contact between the rock <strong>an</strong>d the underlying Athi Series (Saggerson,<br />

1991).


26<br />

The Nairobi phonolite consists <strong>of</strong> a number <strong>of</strong> flows. This is evidenced by age determinations<br />

made on feldspars from the rock, <strong>an</strong>d which r<strong>an</strong>ge from 5,2 to 10,2 Ma, suggesting the<br />

possibility <strong>of</strong> more th<strong>an</strong> one age for the lava. Observations in quarries have shown the various<br />

flows to be generally between 30-46 m thick. Further evidence is provided by borehole<br />

records, which show the thickness <strong>of</strong> the rock unit to be 79 m (BH 2023) at Embakasi in the<br />

south –eastern part <strong>of</strong> the area; <strong>an</strong>d 65 m (BH C201) at Wilson Airport in the western part <strong>of</strong><br />

the area. The base <strong>of</strong> the flow commonly consists <strong>of</strong> pumiceous <strong>an</strong>d dribbly lava overlying a<br />

slightly uneven surface <strong>of</strong> agglomeratic tuff (Saggerson, 1991).<br />

Borehole evidence <strong>an</strong>d flow directions recorded in the field suggest the lavas <strong>of</strong> Nairobi<br />

phonolite having derived from the Muguga area, to the west <strong>of</strong> Nairobi city <strong>an</strong>d present area.<br />

The lavas flowed in a south-easterly stream across Nairobi <strong>an</strong>d the present area, before<br />

consolidation. Maximum thicknesses <strong>of</strong> the lavas therefore occur at/ <strong>an</strong>d around the Jomo<br />

Kenyatta International Airport located in the eastern part <strong>of</strong> the study area. The lavas<br />

generally thin out north-westwards, as well as farther south-eastwards <strong>an</strong>d southwards up to<br />

Athi River township, located beyond <strong>an</strong>d outside the present area (Saggerson, 1991).<br />

Aerial photographs <strong>of</strong> the plain at Jomo Kenyatta International Airport reveal sub-circular<br />

patches with crudely linear arr<strong>an</strong>gements beneath the black cotton soil cover. These patches<br />

probably represent the areas <strong>of</strong> maximum weathering at intersections <strong>of</strong> master joint-pl<strong>an</strong>es<br />

(Scott, 1963).<br />

Parts <strong>of</strong> the Nairobi National Park between Bushy Vale <strong>an</strong>d Rocky Valley show the Nairobi<br />

phonolite separated from the underlying Mbagathi phonolitic trachytes by some thickness <strong>of</strong> a<br />

few feet <strong>of</strong> dark grey tuff, which belongs to the Athi Tuffs <strong>an</strong>d Lake Beds Series. Outcrops<br />

generally reveal minor differences in the Nairobi phonolites. They are characterised by a<br />

marked fissility <strong>an</strong>d platy texture developed locally at <strong>an</strong>d/ or near the surface, <strong>an</strong>d which is<br />

attributed to flow phenomena. They also reveal the vesicular character <strong>of</strong> the upper portion <strong>of</strong><br />

a number <strong>of</strong> flows. However, the rocks are seldom amygdaloidal (Saggerson, 1991).<br />

Fresh h<strong>an</strong>d specimens <strong>of</strong> the rock are black or blue-grey to blue in colour. However, deep<br />

weathering has altered much <strong>of</strong> the upper portions <strong>of</strong> the rock located near the ground surface.<br />

The rock is also lacking in large phenocrysts, the phonolite having undoubtedly resulted from<br />

the rapid cooling <strong>of</strong> <strong>an</strong> already viscous lava.<br />

Weathering usually converts the Nairobi phonolites to pale brown <strong>an</strong>d reddish ferricrete<br />

deposits (murram), which are observed to overly the rocks in much <strong>of</strong> the study area. In other<br />

parts, however, especially in the eastern <strong>an</strong>d south-eastern sections <strong>of</strong> the present area, the<br />

lava surface has been autobrecciated to form a boulder-bed consisting <strong>of</strong> rounded <strong>an</strong>d <strong>an</strong>gular<br />

fragments <strong>of</strong> phonolite cemented in a light brown, weathered base <strong>of</strong> the same material<br />

(Saggerson, 1991).<br />

Thin sections <strong>of</strong> Nairobi phonolite exhibit long tabular phenocrysts <strong>of</strong> s<strong>an</strong>idine in a finegrained<br />

groundmass. The groundmass consists <strong>of</strong> patches <strong>of</strong> soda-amphibole which include<br />

cossyrite <strong>an</strong>d kataphorite, while bright green aegirine <strong>an</strong>d/ or aegirine-augite are found<br />

associated. Nepheline is found occurring as small micro-phenocrysts, or as crystals in the<br />

groundmass; larger crystals <strong>of</strong> nepheline commonly alter into yellowish zeolite, sodalite,<br />

natrolite, <strong>an</strong>alcime <strong>an</strong>d calcite. Strongly pleochroic biotite with magnetite inclusions, or<strong>an</strong>gebrown<br />

mica flakes as well as occasional zeolite-filled vesicles may also be found (Fairburn,<br />

1963). The matrix composition is also summarised in Table (3.4). Table 3.2 gives the results<br />

<strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> Nairobi phonolite, as well as other rocks for comparison purposes.


27<br />

Table 3.2 Chemical <strong>an</strong>alyses <strong>of</strong> strongly alkaline lavas <strong>of</strong> Nairobi phonolite<br />

<strong>an</strong>d other rocks.<br />

186 157 135 132 118 W64 112 99 108<br />

SiO2 38,80 45,65 49,62 51,59 54,81 56,56 57,05 57,94 59,14<br />

Al2O3 14,22 16,12 21,23 14,48 19,10 16,77 19,73 17,00 15,75<br />

Fe2O3 7,28 3,07 3,77 7,77 1,57 3,17 1,85 4,03 6,65<br />

FeO 6,34 9,08 2,02 - 3,52 2,74 4,31 1,94 1,35<br />

MgO 5,65 7,38 0,65 2,90 0,96 1,36 0,34 1,22 0,56<br />

CaO 12,35 9,70 3,40 4,53 1,81 1,18 1,56 0,99 1,42<br />

Na2O 5,00 2,74 9,00 - 6,88 7,53 7,54 7,63 7,23<br />

K2O 2,19 1,42 5,45 - 5,41 5,69 4,85 5,13 4,66<br />

H2O+ 1,80 0,33 1,81 3,75 3,63 1,70 0,54 1,15 1,78<br />

H2O- 1,40 0,68 1,75 3,75 0,57 1,55 0,68 0,94 0,25<br />

TiO2 4,30 3,30 0,83 0,72 0,82 1,07 0,82 1,08 0,50<br />

P2O5 0,72 0,43 0,11 - 0,22 0,13 0,23 0,18 0,09<br />

MnO 0,14 0,13 0,18 - 0,27 0,33 0,35 0,30 0,35<br />

CO2 - - - - 0,14 0,05 - 0,09 0,11<br />

SO3 - - - - - - Tr - -<br />

Cl - - 0,21 - 0,01 - - 0,13 0,02<br />

F - - - - 0,09 - - 0,11 -<br />

S - - Nil - 0,02 - - 0,02 -<br />

Loss on - - - 1,23 - - - - -<br />

ignition<br />

-O eqv - - 0,05 - 0,04 - - 0,08 0,00<br />

Total 100,10 100,03 99,88 88,97 99,79 99,83 99,85 99,80 99,86<br />

186. Nephelinite (augitite). West <strong>of</strong> Ngong Bazaar (W. C. Smith, 1931).<br />

157. Alkali basalt (bas<strong>an</strong>ite). Nyeri road, Aberdare r<strong>an</strong>ge (W. C. Smith, 1931).<br />

135. Phonolite. Ol Esakut (Tilley, 1956).<br />

132. Nairobi phonolite. Nairobi (Analysed by E.A.I.R.B.).<br />

118. Phonolite (Kapiti-type). Nairobi-Nyeri road, Nairobi (Bowen, 1937).<br />

W64. Nairobi phonolite. Nairobi (Analysed by H. LIoyd).<br />

112. Fayalite-bearing phonolite. Nyeri hill (W. C. Smith, 1931).<br />

99. Phonolitic trachyte. Athi plains (Bowen, 1937).<br />

108. Phonolite (Kenya-type). North slope, Suswa (McCall, 1964).<br />

3.3.3 Nairobi trachytes<br />

The Nairobi trachytes exhibit a wide distribution, extending from the Dagoretti – Karen area<br />

located to the west <strong>of</strong> the study area, <strong>an</strong>d eastwards to underly Nairobi city <strong>an</strong>d the present<br />

area. They also extend farther northwards, beyond the present area, to underly areas <strong>of</strong><br />

Kiambu <strong>an</strong>d south Githunguri ( Figure 1.1).<br />

A stratigraphic relationship is revealed by a number <strong>of</strong> borehole records <strong>an</strong>d the m<strong>an</strong>y<br />

outcrops in stream courses found in the immediate vicinity <strong>of</strong> Nairobi city.The Nairobi<br />

trachytes are shown to underly tuffs <strong>an</strong>d a number <strong>of</strong> other trachytes, while they overly the<br />

Kiambu trachyte, Mbagathi phonolitic trachyte <strong>an</strong>d Nairobi phonolite (Saggerson, 1991).<br />

Outcrops <strong>of</strong> the rock are particularly observable in the south-western part <strong>of</strong> the present area,<br />

where the trachytes form a flat-topped plateau terminating in a low but prominent escarpment


28<br />

overlooking the Nairobi National Park. Outcrops <strong>of</strong> individual flows are also represented by<br />

small steps in the escarpment immediately to the west <strong>of</strong>/ <strong>an</strong>d overlooking Nairobi city <strong>an</strong>d<br />

the present study area (Saggerson, 1991).<br />

Borehole records show penetrated rocks as generally occurring in a continous sequence <strong>of</strong><br />

lava flows, varying in thickness from 91 m around Nairobi city centre to 61 m at Ruaraka (to<br />

the north <strong>of</strong> present study area). Records also show the flows as having vesicular upper<br />

contact surfaces <strong>an</strong>d/ or to be separated from one <strong>an</strong>other by sediments or tuffs (BH 3001,<br />

Mbagathi). In addition, stratigraphic information obtained from boreholes <strong>an</strong>d exposures in<br />

quarries in the Nairobi municipality <strong>an</strong>d the present study area, show the Nairobi trachyte<br />

separated from the underlying Nairobi phonolite by a few feet thickness <strong>of</strong> agglomeratic tuff<br />

(Saggerson, 1991).<br />

Table 3.3 Chemical <strong>an</strong>alyses <strong>of</strong> mildly alkaline lavas <strong>of</strong> Nairobi trachyte & other volc<strong>an</strong>ic rocks.<br />

170 151 203 95 W15 87 202<br />

SiO2 43,50 46,74 57,42 59,38 62,56 62,54 59,70<br />

Al2O3 18,65 18,88 17,92 10,84 14,96 15,85 19,20<br />

Fe2O3 1,70 3,11 7,50 3,92 3,25 3,66 8,00<br />

FeO 10,48 7,15 7,50 4,28 3,36 2,62 8,00<br />

MgO 7,42 3,16 1,17 0,89 0,36 0,36 0,30<br />

CaO 10,26 9,30 Tr 3,31 1,21 1,46 2,40<br />

Na2O 2,63 4,17 - 5,60 6,42 6,12 6,40<br />

K2O 0,94 1,67 - 5,01 5,64 5,07 2,80<br />

H2O+ 0,54 0,63 - 3,14 0,54 0,59 -<br />

H2O- 0,80 1,07 - 3,14 0,73 0,45 -<br />

TiO2 2,40 2,44 0,60 0,81 0,66 1,07 1,2<br />

P2O5 0,62 1,29 Nd 1,00 0,10 0,13 -<br />

MnO 0,12 0,18 - 0,35 0,19 - -<br />

CO2 - 0,16 3,42 1,30 Nil - -<br />

SO3 - - 0,15 0,04 0,00 0,08 -<br />

Cl - - - 0,01 0,04 0,04 -<br />

F - 0,11 - - - - -<br />

S - - - - - - -<br />

-O eqv - 0,05 - - 0,02 0,01 -<br />

Totals 100,06 100,01 - 99,88 100,00 100,03 100,00<br />

170. Alkali basalt, Rogati river, Kikuyu ( W. C. Smith, 1931).<br />

151. Olivine basalt, Kijabe Hill (Sh<strong>an</strong>d, 1937).<br />

203. Tuff, junction Ngong-Dagoretti roads, Impala Institute, Nairobi (W. C. Smith, 1931).<br />

95. Alkali trachyte (Plateau Trachyte Series), Magadi Hospital (Baker, 1958).<br />

202. Welded tuff/ Nairobi building stone (E.A.I.R.B).<br />

87. Trachyte phonolitique (soda-trachyte), Kikuyu escarpment (Lacroix, 1923).<br />

W15. Nairobi trachyte, Kenyatta Avenue, Nairobi (E.A.I.R.B).<br />

H<strong>an</strong>d specimens <strong>of</strong> Nairobi trachyte are greenish-grey in colour, <strong>an</strong>d occasionally porphyritic<br />

exhibiting tabular phenocrysts <strong>of</strong> feldspar in a predomin<strong>an</strong>tly fine-grained groundmass. Dark<br />

streaks, patches <strong>an</strong>d b<strong>an</strong>ds may also be found <strong>an</strong>d serve to indicate segregation <strong>of</strong> the mafic<br />

constituents. The more homogeneous varieties <strong>of</strong> the rock also show a trachytic texture<br />

characterised by a typical silvery lustre. Lamination <strong>an</strong>d b<strong>an</strong>ding are also characteristic <strong>an</strong>d<br />

are a result <strong>of</strong> flow <strong>an</strong>d flattening within the lava body. The flow is further emphasised by<br />

fluxionally arr<strong>an</strong>ged feldspar laths, as well as variations in rock composition <strong>an</strong>d colour


29<br />

caused by partial segregation <strong>of</strong> the dark <strong>an</strong>d light-coloured constituents present, during<br />

crystallisation (Saggerson, 1991).<br />

Thin sections show small tabular s<strong>an</strong>idine phenocrysts (exhibiting Carlsbad twinning) set in a<br />

fine-grained groundmass. The groundmass shows a marked trachytic texture <strong>an</strong>d consists <strong>of</strong><br />

orthoclase associated with soda-amphiboles <strong>an</strong>d pyroxenes, as well as rare iron ore. The sodarich<br />

minerals consist <strong>of</strong> deep brown cossyrite <strong>an</strong>d green aegirine-augite, <strong>an</strong>d occur as<br />

prominent moss-like aggregates <strong>of</strong> prismatic crystals. The alternating brown <strong>an</strong>d green<br />

patches serve to impart a mottled appear<strong>an</strong>ce to the rock in h<strong>an</strong>d specimens.<br />

Table 3.3 gives results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> Nairobi trachyte, as well as other rocks for<br />

comparison purposes. Results <strong>of</strong> mineralogical <strong>an</strong>alyses <strong>of</strong> Nairobi trachyte, <strong>an</strong>d a number <strong>of</strong><br />

other rocks underlying Nairobi <strong>an</strong>d the present study area, are summarised in Table 3.4<br />

(Saggerson, 1991).<br />

Table 3.4 Mineral compositions <strong>of</strong> trachytes <strong>an</strong>d phonolites from Nairobi <strong>an</strong>d the study area<br />

(after Saggerson, 1991).<br />

Rock type:<br />

Nairobi Kiambu Nairobi K<strong>an</strong>dizi Mbagathi Kapiti<br />

trachyte trachyte phonolite phonolite phonolitic phonolite<br />

trachyte<br />

S<strong>an</strong>idine X X R Ac<br />

Anorthoclase R X X<br />

Nepheline<br />

X<br />

Tit<strong>an</strong>augite R X<br />

Olivine<br />

R<br />

Phenocrysts<br />

Matrix<br />

Biotite<br />

X<br />

Alkali feldspar X X X X X X<br />

Nepheline X X X X<br />

Quartz<br />

Aegirine<br />

Aegirine-augite<br />

Ok X X On X X<br />

Cossyrite X X X X X<br />

α-kataphorite X X I X X<br />

β-kataphorite<br />

X<br />

Riebeckite<br />

X<br />

Arfvedsonite<br />

Biotite<br />

X<br />

Barkevikite<br />

R<br />

α-kataphorite – colourless to pale brownish yellow, smoky-brown to rose-red <strong>an</strong>d/ or<br />

reddish yellow<br />

- cleavage 124°<br />

β-kataphorite – pale smoky-brown to greenish yellow, deep purplish brown to opaque<br />

X - present; R - rare; A - altered; Ac – accessory<br />

Ok – overgrowths on kataphorite<br />

On – overgrowths on nepheline<br />

I – including large crystals


30<br />

Chapter 4<br />

Field methods<br />

4.1 Introduction<br />

The principal objectives <strong>of</strong> the field study were to:<br />

(i)<br />

(ii)<br />

(iii)<br />

(iv)<br />

Provide information on the nature <strong>an</strong>d distribution <strong>of</strong> black cotton soils (exp<strong>an</strong>sive<br />

<strong>an</strong>d reactive soils) <strong>an</strong>d red soils <strong>of</strong> the study area.<br />

Acquire site specific information on environmental characteristics which impact<br />

on in situ soil behaviour (groundwater conditions, climate, vegetation).<br />

Investigate <strong>an</strong>d model depth variation <strong>of</strong> soils across the study area.<br />

Compare in situ soil properties, behaviour <strong>an</strong>d derived parameters with those<br />

predicted <strong>an</strong>d/ or derived from laboratory studies.<br />

The cone penetration sounding, soil augering, in-situ description <strong>of</strong> soils <strong>an</strong>d underlying rock,<br />

tests for possible carbonate content, as well as undisturbed <strong>an</strong>d disturbed soil sampling were<br />

carried out, among others. Documentation <strong>of</strong> results was done according to known st<strong>an</strong>dards<br />

(BGR, 1996; DIN 4021/ 4096, 1980; IAEG, 1981; ISRM, 1978; NLfB & BGR, 1991).<br />

The sampling technique was selected <strong>an</strong>d pl<strong>an</strong>ned in such a way as to facilitate collection <strong>an</strong>d<br />

provision <strong>of</strong> high quality <strong>an</strong>d detailed data for validating models <strong>an</strong>d assessing variation <strong>of</strong><br />

<strong>engineering</strong> soil properties across the project area.<br />

4.2 Investigation <strong>an</strong>d sampling scheme<br />

A number <strong>of</strong> site investigation <strong>an</strong>d sampling schemes have been devised by practical<br />

statistici<strong>an</strong>s, <strong>an</strong>d are available for use by geologists (Koch <strong>an</strong>d Link, 1970; Krumbein <strong>an</strong>d<br />

Graybill, 1965; Meyer, 1975; Wetherill, 1966; ). According to Cheeney (1983), the principal<br />

motivation for the diversity <strong>of</strong> available procedures is the need to achieve maximum<br />

efficiency <strong>an</strong>d precision within constraints <strong>of</strong> cost, time, effort, available facilities, m<strong>an</strong>power<br />

skills <strong>an</strong>d specimen accessibility. However, the application <strong>of</strong> these schemes may be<br />

hampered <strong>an</strong>d/ or limited by the fact that, <strong>geological</strong> materials <strong>an</strong>d observations are frequently<br />

hard won under difficult circumst<strong>an</strong>ces.<br />

A site investigation/ sampling pl<strong>an</strong> is a fundamental part <strong>of</strong> the study <strong>of</strong> soils, especially as<br />

may regard the assessment, <strong>an</strong>alysis <strong>an</strong>d evaluation <strong>of</strong> the spatial distribution <strong>of</strong> import<strong>an</strong>t<br />

<strong>engineering</strong> <strong>geological</strong> characteristics <strong>an</strong>d/ or parameters. The use <strong>of</strong> <strong>an</strong> unsuitable <strong>an</strong>d<br />

inadequate procedure could lead to bias in the observation <strong>an</strong>d selection <strong>of</strong> specimens, with a<br />

consequent distortion <strong>of</strong> the outcome <strong>of</strong> import<strong>an</strong>t soil studies <strong>an</strong>d evaluations such as<br />

hypothesis testing <strong>an</strong>d statistical tests. It is therefore essential that the field investigation/<br />

sampling method adopted is such as to allow for the acquisition <strong>of</strong> necessary information <strong>an</strong>d<br />

soil samples which are free as much as possible from bias. This is because a biassed set <strong>of</strong><br />

observations <strong>an</strong>d/ or sample would only serve to give a biassed impression <strong>of</strong> the parent<br />

population from which it was drawn (Cochr<strong>an</strong>, 1977).<br />

In the present study, field investigations <strong>an</strong>d sampling were carried out with the aim <strong>of</strong><br />

achieving both descriptive <strong>an</strong>d <strong>an</strong>alytical objectives <strong>of</strong> soils. Among the <strong>an</strong>alytical objectives<br />

is hypothesis testing, the methods <strong>of</strong> which are based on samples. Attempts were therefore<br />

made to avoid choosing <strong>an</strong>d applying defective site investigation <strong>an</strong>d sampling practices that


32<br />

could otherwise lead to erroneous decisions regarding assessments <strong>an</strong>d estimation <strong>of</strong><br />

parameters <strong>of</strong> parent populations (<strong>of</strong> selected soil types). The boundaries/ limits <strong>of</strong> the areas to<br />

be site investigated <strong>an</strong>d sampled were clearly defined <strong>an</strong>d marked. The investigation <strong>an</strong>d<br />

sampling points were then arr<strong>an</strong>ged <strong>an</strong>d distributed in such a way that they covered the<br />

population (or projected area <strong>of</strong> interest) more or less uniformly.<br />

A systematic site investigation <strong>an</strong>d/ or sampling technique was employed in the current study<br />

as the st<strong>an</strong>dard procedure for the black <strong>clays</strong> (Fig. 4.1). The procedure generally involves<br />

dividing the population or the projected area <strong>of</strong> investigation <strong>an</strong>d sampling into a number <strong>of</strong><br />

mutually exclusive subpopulations or strata, over the spatial r<strong>an</strong>ge <strong>of</strong> interest. In this way,<br />

there would be as m<strong>an</strong>y strata as there are to be specimens in the projected sample. During<br />

site investigation <strong>an</strong>d sampling, the position <strong>of</strong> the first specimen taken from the first stratum,<br />

is decided at r<strong>an</strong>dom. Succeeding specimens would then be taken from the same/ similar<br />

position but in their respective strata. According to Cheeney (1983), systematic investigation<br />

<strong>an</strong>d sampling is straightforward, easy to execute, fast; <strong>an</strong>d involves relatively low labour<br />

input. In addition, this site investigation/ sampling scheme is usually adopted as <strong>an</strong> effort<br />

towards improving on the precision <strong>of</strong> estimation <strong>of</strong> some population parameter.<br />

Figure 4.1 illustrates the field arr<strong>an</strong>gement <strong>of</strong> systematic site investigation <strong>an</strong>d sampling as<br />

effected for the black <strong>clays</strong> in this study. Site investigation (soil augering & cone penetration)<br />

points were located at equal intervals <strong>of</strong> 250 m dist<strong>an</strong>ce along pl<strong>an</strong>ned field pr<strong>of</strong>iles. A total<br />

<strong>of</strong> five parallel field pr<strong>of</strong>iles (A, B, C, D, E) in <strong>an</strong> east-west direction, as well as <strong>an</strong>other five<br />

similar pr<strong>of</strong>iles (F, G, H, I, J) in a north-south direction; were used over the extent <strong>of</strong><br />

investigation interest <strong>of</strong> black <strong>clays</strong> across the study area. The interpr<strong>of</strong>ile separation dist<strong>an</strong>ce<br />

was 1000 m for the east-west running pr<strong>of</strong>iles <strong>an</strong>d 2000 m for the north-south pr<strong>of</strong>iles.<br />

Disturbed <strong>an</strong>d undisturbed soil sampling were effected in h<strong>an</strong>d-dug excavation pits (Plates 4.1<br />

& 4.2).<br />

Plates 4.1(a) & (b) H<strong>an</strong>d-dug excavation pits <strong>an</strong>d undisturbed sampling in black <strong>clays</strong>.


33<br />

Plates 4.2(a) & (b) H<strong>an</strong>d-dug excavation pits <strong>an</strong>d undisturbed sampling in red soils.<br />

The nature <strong>an</strong>d depth <strong>of</strong> soils between adjacent excavation pits <strong>an</strong>d/ or sampling points was<br />

investigated <strong>an</strong>d confirmed by employing soil augering (Plate 4.3) <strong>an</strong>d cone penetration<br />

sounding (Plate 4.4) techniques.<br />

Plate 4.3 Augering <strong>of</strong> red soils.<br />

Plate 4.4 Cone penetration sounding <strong>of</strong> soil<br />

depths.


34<br />

The red soils in this study occur mainly within the built-up areas <strong>of</strong> Nairobi metropolit<strong>an</strong>.<br />

Extensively open grounds to facilitate more systematic site investigation <strong>an</strong>d sampling were<br />

therefore hard to come by. As a result, only r<strong>an</strong>dom site studies <strong>an</strong>d sampling <strong>of</strong> red soils<br />

were effected in this study. However, in <strong>an</strong> effort to achieve precise assessments in the<br />

variations <strong>of</strong> soil depths <strong>an</strong>d <strong>engineering</strong> properties within these soils, results obtained from<br />

current investigation have been supplemented with some results <strong>of</strong> previous works (Ot<strong>an</strong>do,<br />

1996) covering the same aspects; <strong>an</strong>d this especially in the western <strong>an</strong>d northern sections <strong>of</strong><br />

the project area.<br />

4.3 Cone penetration depth sounding<br />

Cone penetration sounding technique was used to determine soil depths across the project<br />

area. A combined presentation <strong>of</strong> the sounding pl<strong>an</strong> <strong>an</strong>d obtained results for both red soils<br />

<strong>an</strong>d black <strong>clays</strong> is shown in Fig. (4.2) in which values at site investigation points refer to<br />

respective depths. The red soils are mainly confined to the central-western <strong>an</strong>d north-western<br />

parts <strong>of</strong> the project area, <strong>an</strong>d show depths <strong>of</strong> just over 4m to 18m. The black <strong>clays</strong> cover the<br />

rest <strong>of</strong> the project area <strong>an</strong>d commonly vary in their depths from about 1m to around 4m. In<br />

general, however, both red soils <strong>an</strong>d black <strong>clays</strong> show progressively decreased soil depths<br />

eastwards <strong>an</strong>d south-eastwards across the study area, just as indicated by the direction <strong>of</strong><br />

arrows in Fig. (4.2).<br />

Depth variations <strong>of</strong> red soils <strong>an</strong>d black <strong>clays</strong> are also summarised in contour form (Fig. 4.3),<br />

block diagram (Fig. 4.4), as well as contour overlays on a base map (Appendix D: Fig. D1).


35<br />

36°58'E<br />

1°24'S<br />

North<br />

18 17 16 14 13 12 11 1 1 1 1 1 1 4 4<br />

16 15 14 13 12 4 4 4<br />

1 1 1 1 1 1 1<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000 22000<br />

15 14 13 12 11 1 1 1 1 2 2 11 1 11 1 11 1 11 1<br />

14 14 12 1 1 1 2 2 2 1 1 1 1 1 1 1 1 1<br />

13 12 11 1 1 1 4 2<br />

3322<br />

332 21 1 1 1 1 1 1 1 1 1 1<br />

12 11 10 1 1 1 1 1 1 1 1 1 1 331 1 1 1 1 1<br />

11 10 9 1 1 2 2 1 1 1 1 1 1 1 1<br />

10 9 8 8 1 1 2 2 2 1 1 1 1 1 1 3331 1 1 1<br />

9 9 8 7 2 2 1 1 11 3332 1 1 1 1 1 1 1 1<br />

8 8 8 0 1 2 2 0 1 3 1 1 1 1 2 3 1 1 1 1<br />

6 5 6 5 1 1 1 2 1 1 1 2 1 1 1 1 1<br />

1 0 0 0 1 1 0 1 1 1 1 2 1 1 1<br />

1 2 2 2 1 0 1 0 2 2 2 0 1 1 0 1 1 0 1 1<br />

2 2 1 0 1<br />

2 1 2 0 0 1 0 0 0 0<br />

Dist<strong>an</strong>ce (m)<br />

Figure 4.2. Depth sounding results showing values in metres.<br />

15000<br />

East<br />

1°15'S<br />

36°45'E<br />

10000<br />

5000<br />

0<br />

Dist<strong>an</strong>ce (m)


36<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Black <strong>clays</strong> Red soils<br />

36°58'E<br />

1°24'S<br />

1°15'S<br />

15000<br />

Dist<strong>an</strong>ce (m)<br />

10000<br />

36°45'E<br />

West<br />

5000<br />

0<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000 22000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Figure 4.3. Depth variations in red soils <strong>an</strong>d black <strong>clays</strong>.


4<br />

3<br />

2<br />

1<br />

0<br />

37<br />

Depth (m)<br />

15<br />

10<br />

5<br />

1°15'S<br />

Dist<strong>an</strong>ce (m)<br />

36°58'E<br />

West<br />

1°24'S<br />

36°45'E<br />

5<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

South Dist<strong>an</strong>ce (m)<br />

Black <strong>clays</strong> Red soils<br />

Depth (m)<br />

Figure 4.4. Block diagram showing depth variations in red soils <strong>an</strong>d black <strong>clays</strong>.


38<br />

4.4 V<strong>an</strong>e test<br />

4.4.1 Introduction<br />

The field v<strong>an</strong>e allows for the in situ measurement <strong>of</strong> the shear strength <strong>of</strong> very s<strong>of</strong>t <strong>an</strong>d<br />

sensitive clay soils (up to 20 kPa or less), from which it would be otherwise very difficult to<br />

prepare undisturbed specimens for other types <strong>of</strong> test (Head, 1988; Richwien & Lesny, 2000).<br />

The v<strong>an</strong>e shear tests therefore serve to determine the overall shear strength <strong>of</strong> undrained soils<br />

through a relatively faster shearing process. It is, however, not possible to separate the shear<br />

strength so determined into the components <strong>an</strong>d/ or shear parameters <strong>of</strong> friction <strong>an</strong>d cohesion,<br />

since the effective shear strength in the shear surface is unknown (DIN 4096, 1980). Design<br />

<strong>an</strong>d related calculations involving projected works in s<strong>of</strong>t undrained soils therefore assume <strong>an</strong><br />

<strong>an</strong>gle <strong>of</strong> internal friction (φ) <strong>of</strong> zero.<br />

In the present area, isolated sections <strong>of</strong> very s<strong>of</strong>t soil materials occur in the form <strong>of</strong> peaty <strong>an</strong>d<br />

swampy environs, <strong>an</strong>d/ or areas <strong>of</strong> impeded drainage in those parts <strong>of</strong> black cotton soils; as<br />

well as in some relatively dense forested areas <strong>of</strong> red soils.<br />

4.4.2 Testing procedure<br />

The principle <strong>of</strong> the v<strong>an</strong>e test is shown in Figure 4.2. This study adopted the British St<strong>an</strong>dard<br />

(BS 1377: 1975, Test 18 <strong>an</strong>d/ or DIN 4096: 1980) field v<strong>an</strong>e testing procedure, which<br />

specifies a rate <strong>of</strong> rotation <strong>of</strong> the v<strong>an</strong>e <strong>of</strong> 6-12° /minute. The apparatus was designed to have a<br />

capacity <strong>of</strong> up to 240 Nm in terms <strong>of</strong> applied external torque, thereby capable <strong>of</strong> measuring<br />

shear strength <strong>of</strong> s<strong>of</strong>t to stiff soils <strong>of</strong> up to 100 kN/m², with <strong>an</strong> accuracy <strong>of</strong> 0,2 kN/m².<br />

The apparatus used in this investigation consisted <strong>of</strong> a four-blade cruciform v<strong>an</strong>e <strong>of</strong> diameter<br />

(D) <strong>of</strong> 50 mm <strong>an</strong>d height (H) <strong>of</strong> 100mm, attached to the lower end <strong>of</strong> a rod <strong>of</strong> small diameter<br />

(Plate 4.5). The test involved pushing the cruciform v<strong>an</strong>e into the soil followed by its rotation.<br />

Figure 4.5 Principle <strong>of</strong> v<strong>an</strong>e shear test showing v<strong>an</strong>e blades <strong>an</strong>d cylinder <strong>of</strong> soil rotated by<br />

v<strong>an</strong>e (after Head, 1988).


39<br />

The upper ends <strong>of</strong> the v<strong>an</strong>es were operated at a penetration depth <strong>of</strong> at least 300 mm from the<br />

ground surface, to avoid possibly disturbed upper soil zones (DIN 4096, 1980). The torque<br />

required to cause rotation <strong>of</strong> the cylinder <strong>of</strong> soil enclosing the v<strong>an</strong>e was measured <strong>an</strong>d/ or read<br />

<strong>of</strong>f, <strong>an</strong>d this enabled the undrained shear strength <strong>of</strong> the s<strong>of</strong>t soil to be calculated using the<br />

equation below (DIN 4096, 1980; Head, 1988; Richwien & Lesny, 2000) i.e.,<br />

where c = v<strong>an</strong>e shear strength (kN/m²)<br />

T = applied torque (kN.m)<br />

D = diameter <strong>of</strong> v<strong>an</strong>e (m)<br />

<strong>an</strong>d H = 2D for apparatus used in this study (m)<br />

c = 6T/(7πD³) kN/m² (4.1)<br />

Plates 4.5(a) & (b) V<strong>an</strong>e shear testing in s<strong>of</strong>t black <strong>clays</strong>.<br />

A general guide to the description <strong>of</strong> soils based on results <strong>of</strong> v<strong>an</strong>e shear tests is summarised<br />

in Table 4.1 (BS 1377: 1975).


40<br />

Table 4.1 Classification <strong>of</strong> s<strong>of</strong>t soils based on v<strong>an</strong>e shear strength (BS 1377: 1975).<br />

General descriptive term<br />

Maximum shear stress<br />

for strength<br />

(kN/m²)<br />

Very s<strong>of</strong>t 20<br />

S<strong>of</strong>t 40<br />

S<strong>of</strong>t to firm 60<br />

Firm 90<br />

4.4.3 Results <strong>of</strong> study<br />

Results <strong>of</strong> field v<strong>an</strong>e shear tests performed in this study are summarised in Tables 4.2, 4.3 <strong>an</strong>d<br />

4.4 as shown below. The variations <strong>of</strong> v<strong>an</strong>e shear strength with soil depth are presented in<br />

Figure 4.4.<br />

Table 4.2 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t black <strong>clays</strong> at Madaraka West, Nairobi.<br />

Date: 14.01.2001 Spring = Tr/ θf i.e. K = 1333,33<br />

Location: Madaraka West Const<strong>an</strong>t, K Nmm/deg.<br />

Soil type: Black <strong>clays</strong><br />

Depth, Moisture Bulk Angle <strong>of</strong> Torque, V<strong>an</strong>e shear Soil<br />

Dp (m) Content, density Rotation, θf Tr (Nmm) Strength, strength<br />

Wn (%) (g/cm³) (degrees) c (kPa) description<br />

0,3 40,53 1,56 10,6 14166,67 30,94 S<strong>of</strong>t<br />

0,6 40,46 1,74 14,3 19000,00 41,50 s<strong>of</strong>t to firm<br />

1,0 24,68 1,78 45,6 60833,33 132,86 firm to stiff<br />

1,5 28,23 1,78 60,6 80833,33 176,53 Stiff<br />

The s<strong>of</strong>t <strong>an</strong>d sensitive clay soils are characterised by a very s<strong>of</strong>t to firm consistency, up to<br />

depths <strong>of</strong> 1 m for black <strong>clays</strong> <strong>an</strong>d 3 m for red soils. The black clay vari<strong>an</strong>ts exhibit undrained<br />

v<strong>an</strong>e shear strength <strong>of</strong> 30,94 – 41,50 kPa; while the red soil equivalents have values <strong>of</strong> 10,92<br />

– 87,36 kPa. There is a general increase in the undrained shear strength with depth, in both<br />

soil types. This is probably due to preconsolidation <strong>of</strong> deeper soil horizons by overburden<br />

material above, so that bulk density <strong>of</strong> the soils generally increases with depth (Tables 4.2, 4.3<br />

& 4.4).<br />

Table 4.3 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t red soils at Arboretum, Nairobi.<br />

Date: 11.01.2001 Spring = Tr/ θf i.e. K = 1333,33<br />

Location: Arboretum Const<strong>an</strong>t, K Nmm/deg.<br />

Soil type: Red friable <strong>clays</strong><br />

Depth, Moisture Bulk Angle <strong>of</strong> Torque, V<strong>an</strong>e shear Soil<br />

Dp (m) Content, density Rotation, θf Tr (Nmm) Strength, c strength<br />

Wn (%) (g/cm³) (degrees) (kPa) description<br />

0,3 39,44 1,37 3,8 5000,00 10,92 very s<strong>of</strong>t<br />

1,0 24,03 1,41 30,0 40000,00 87,36 Firm<br />

2,0 24,25 1,41 36,5 48616,47 106,18 firm to stiff<br />

3,0 25,07 1,52 38,3 51020,39 111,43 firm to stiff


41<br />

Table 4.4 Results <strong>of</strong> field v<strong>an</strong>e shear tests in s<strong>of</strong>t red soils at Kenya High School, Nairobi.<br />

Date: 11.01.2001 Spring = Tr/ θf i.e. K = 1333,33<br />

Location: Kenya High Const<strong>an</strong>t, K Nmm/deg.<br />

Soil type: Red friable <strong>clays</strong><br />

Depth, Dp Moisture Bulk Angle <strong>of</strong> Torque, V<strong>an</strong>e shear Soil<br />

(m) Content, density Rotation, θf Tr (Nmm) strength, c strength<br />

Wn (%) (g/cm³) (degrees) (kPa) description<br />

0,3 37,85 1,36 5,7 7555,19 16,50 very s<strong>of</strong>t<br />

1,0 26,86 1,40 15,1 20147,16 44,00 s<strong>of</strong>t to firm<br />

2,0 27,05 1,43 21,4 28558,60 62,37 Firm<br />

3,0 27,78 1,68 22,7 30220,74 66,00 Firm<br />

The variation <strong>of</strong> v<strong>an</strong>e shear strength with soil depth tends to fit closely to a logarithmic<br />

relationship (Fig. 4.6), with relatively strong correlation, i.e.R = 0,94 for black <strong>clays</strong>; <strong>an</strong>d R =<br />

0,97 – 1,0 for red soils.<br />

V<strong>an</strong>e shear strength (kPa)/ depth (m)<br />

200,00<br />

180,00<br />

c = 95,496Ln(Dp) + 126,71<br />

R 2 = 0,8766<br />

V<strong>an</strong>e shear strength c (kPa)<br />

160,00<br />

140,00<br />

120,00<br />

100,00<br />

80,00<br />

60,00<br />

40,00<br />

c = 44,727Ln(Dp) + 72,397;<br />

R 2 = 0,9397<br />

c = 22,33Ln(Dp) + 43,936;<br />

R 2 = 0,9901<br />

Red soil,<br />

Kenya High<br />

Red soil,<br />

Arboretum<br />

Black <strong>clays</strong>,<br />

Madaraka<br />

Logarithmic<br />

(Kenya High)<br />

Logarithmic<br />

(Arboretum)<br />

Logarithmic<br />

(Madaraka)<br />

20,00<br />

0,00<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5<br />

Depth Dp (m)<br />

Figure 4.6 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with soil depth for the s<strong>of</strong>t red friable <strong>clays</strong><br />

<strong>an</strong>d s<strong>of</strong>t black <strong>clays</strong>.


42<br />

The variation <strong>of</strong> v<strong>an</strong>e shear strength with natural moisture content <strong>of</strong> soils closely<br />

approximates a polynomial relationship with strong correlation for both soil types, giving R =<br />

1.0 for black <strong>clays</strong> <strong>an</strong>d R = 0,91 - 0,99 for red soils (Fig. 4.7).<br />

V<strong>an</strong>e shear strength (kPa)/ natural moisture content (%)<br />

200,00<br />

180,00<br />

c = -1,5051Wn 2 + 91,987Wn - 1220,7<br />

R 2 = 0,9976<br />

160,00<br />

140,00<br />

Red soil, Kenya<br />

High<br />

V<strong>an</strong>e shear strength c (kPa)<br />

120,00<br />

100,00<br />

80,00<br />

c = -1,6845Wn 2 + 101,55Wn - 1373,9<br />

R 2 = 0,9834<br />

Red soil,<br />

Arboretum<br />

Black <strong>clays</strong>,<br />

Madaraka<br />

Polynomial<br />

(Madaraka)<br />

Polynomial<br />

(Kenya High)<br />

60,00<br />

c = 0,3359Wn^2 - 24,801Wn + 473,93<br />

R^2 = 0,8292<br />

40,00<br />

20,00<br />

0,00<br />

0 5 10 15 20 25 30 35 40 45<br />

Natural moisture content Wn (%)<br />

Figure 4.7 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with natural moisture content <strong>of</strong> soils for the<br />

s<strong>of</strong>t red friable <strong>clays</strong> <strong>an</strong>d s<strong>of</strong>t black <strong>clays</strong>.<br />

V<strong>an</strong>e shear strength <strong>an</strong>d bulk density were observed to generally increase with increased soil<br />

depths (Tables 4.2, 4.3, 4.4). So far, the variation <strong>of</strong> v<strong>an</strong>e shear strength with bulk density is<br />

best described by a polynomial relationship (Fig. 4.8) with a very strong correlation (R = 0,97<br />

for black <strong>clays</strong>; R = 0,99-1,0 for red soils).


43<br />

V<strong>an</strong>e shear strength (kPa)/ bulk density (g/cm³)<br />

200,00<br />

c = 12597ρ 2 - 41512ρ + 34133<br />

180,00<br />

R 2 = 0,9366<br />

V<strong>an</strong>e shear strength c (kPa)<br />

160,00<br />

140,00<br />

120,00<br />

100,00<br />

80,00<br />

60,00<br />

40,00<br />

c = -4406,1ρ 2 + 13695ρ - 10461<br />

R 2 = 0,9714<br />

c = -3406,9ρ 2 + 10178ρ - 7533,1<br />

R 2 = 0,9996<br />

Kenya High<br />

School<br />

Arboretum<br />

Madaraka West<br />

Polynomial<br />

(Madaraka West)<br />

Polynomial<br />

(Arboretum)<br />

Polynomial<br />

(Kenya High<br />

School)<br />

20,00<br />

0,00<br />

1,2 1,4 1,6 1,8 2<br />

Bulk density ρ (g/cm³)<br />

Figure 4.8 Variation <strong>of</strong> field v<strong>an</strong>e shear strength with bulk density <strong>of</strong> soils for the<br />

s<strong>of</strong>t red friable <strong>clays</strong> <strong>an</strong>d s<strong>of</strong>t black <strong>clays</strong>.<br />

4.4.4 Applications<br />

In practice, determination <strong>an</strong>d knowledge <strong>of</strong> the values <strong>of</strong> low shear strength <strong>of</strong> s<strong>of</strong>t <strong>an</strong>d/ or<br />

weak soils to a reasonable degree <strong>of</strong> accuracy is essential, <strong>an</strong>d some <strong>of</strong> the applications are as<br />

mentioned below.<br />

First <strong>of</strong> all, the determined low shear strength could be useful in indicating the maximum safe<br />

bearing pressure which a stratum <strong>of</strong> very s<strong>of</strong>t soil could sustain initially, when external loads<br />

<strong>of</strong> constructed <strong>engineering</strong> structures are placed <strong>an</strong>d/ or imposed on them. This would in turn<br />

determine <strong>an</strong>d influence the designed thickness <strong>of</strong> such structures (emb<strong>an</strong>kments, etc.) which<br />

could be built <strong>an</strong>d placed as a first stage. Thereafter, subsequent consolidation would serve to<br />

increase the shear strength <strong>of</strong> the soil so that construction could proceed in stages, based on<br />

the shear strength criterion (Head, 1988).<br />

Secondly, the shear strength <strong>of</strong> s<strong>of</strong>t strata near the ground surface is usually used for the<br />

estimation <strong>of</strong> “negative skin friction” resulting from adhesion on piles driven down to a firmer<br />

stratum below (Lambe <strong>an</strong>d Whitm<strong>an</strong>, 1979). Infact, s<strong>of</strong>t soil deposits which have been<br />

extensively disturbed by the effect <strong>of</strong> driving <strong>of</strong> piles commonly end up exhibiting signific<strong>an</strong>t<br />

amounts <strong>of</strong> remoulded strength.


44<br />

And thirdly, the field v<strong>an</strong>e apparatus c<strong>an</strong> be used to investigate <strong>an</strong>d provide the relationship<br />

between undrained shear strength <strong>an</strong>d moisture content in cohesive soils; <strong>an</strong>d/ or the<br />

relationship between shear strength <strong>an</strong>d soil moisture suction (Lewis <strong>an</strong>d Ross, 1955).


45<br />

Chapter 5<br />

Types <strong>of</strong> soil<br />

5.1 Superficial deposits<br />

Superficial deposits <strong>of</strong> Recent age consist <strong>of</strong> alluvia <strong>an</strong>d conglomerates exposed in the<br />

principal river courses <strong>an</strong>d their tributaries, especially in the southern <strong>an</strong>d south-eastern<br />

sections <strong>of</strong> the study area. Accumulations <strong>of</strong> alluvium occur mainly in the form <strong>of</strong> old river<br />

sediments. Also included are new sediments currently being added to the flood plain without<br />

developed morphology but have a more humic surface horizon. On the Athi plains, these soils<br />

have also developed from the drying up <strong>of</strong> papyrus swamps, either as a result <strong>of</strong> c<strong>an</strong>alisation<br />

<strong>of</strong> the rivers found <strong>an</strong>d/ or drier climatic conditions. Exposures <strong>of</strong> a conglomerate b<strong>an</strong>d<br />

composed <strong>of</strong> boulders <strong>an</strong>d pebbles <strong>of</strong> local lava occur on the north b<strong>an</strong>k <strong>of</strong> the Mbagathi river,<br />

within the Nairobi National Park. The boulders <strong>an</strong>d pebbles are partly indurated while, on the<br />

other h<strong>an</strong>d, the more modern gravels also found exposed along the course <strong>of</strong> the Mbagathi<br />

river are largely uncemented.<br />

Deposits <strong>of</strong> shallow stony soils associated with rock outcrops characterise the courses <strong>of</strong> most<br />

<strong>of</strong> the streams draining the high ground east <strong>of</strong> the Rift, especially in the northern <strong>an</strong>d northwestern<br />

sections <strong>of</strong> the present area. The soils are variously developed <strong>an</strong>d originated from<br />

steep slope areas where they derived by processes <strong>of</strong> accelerated erosion <strong>of</strong> the various soils<br />

found, <strong>an</strong>d which have lost their original characteristics (Saggerson, 1991). They are very<br />

shallow <strong>an</strong>d occur on the main valley sides in the form <strong>of</strong> pockets on slight shelves <strong>an</strong>d<br />

between boulders.<br />

Isolated areas <strong>of</strong> impeded drainage also occur across the study area, <strong>an</strong>d are characterised by<br />

development <strong>of</strong> swamps with black peaty soils as well as dark greyish-brown mottled <strong>clays</strong>,<br />

both <strong>of</strong> which form accumulations <strong>of</strong> poorly-drained soils. The peaty swamps constitute a<br />

very minor soil group in the study area <strong>an</strong>d usually occur in the form <strong>of</strong> narrow stretches<br />

confined to the b<strong>an</strong>ks <strong>of</strong> streams <strong>an</strong>d/ or rivers flowing through them. The peaty soils found<br />

associated are characterised by the growth <strong>of</strong> papyrus pl<strong>an</strong>ts <strong>an</strong>d a typical grass vegetation<br />

cover. These soils also exhibit a high humus content <strong>an</strong>d are usually subjected to seasonal or<br />

perm<strong>an</strong>ent water table. They are usually covered by water during flood times (Saggerson,<br />

1991). The swampy areas clearly show up as dark tones on aerial photographs.<br />

The mottled <strong>clays</strong> commonly exhibit a dark grey to greyish-brown humic (2-3% carbon)<br />

topsoil overlying a greyish-brown <strong>an</strong>gular blocky mottled subsoil. These ill-drained soils<br />

generally occur in the main valley bottoms <strong>of</strong> old drainage lines <strong>an</strong>d/ or depressions where the<br />

greyish-brown mottled <strong>clays</strong> commonly fill in areas <strong>of</strong> fault troughs <strong>an</strong>d/ or occupy the flat<br />

ground at the foot <strong>of</strong> some fault scarps where drainage is locally impeded. The depressions<br />

are most probably a result <strong>of</strong> minor fissuring which caused the formation <strong>of</strong> sinkholes at the<br />

junctions <strong>of</strong> such fissures. Later on, the sinkholes <strong>an</strong>d old drainage lines could have been<br />

filled in with deposited materials, which also acted as the parent material for the development<br />

<strong>an</strong>d formation <strong>of</strong> the mottled <strong>clays</strong> under conditions <strong>of</strong> impeded drainage. According to Scott<br />

(1963), the mottled <strong>clays</strong> most probably owe their origin <strong>an</strong>d development from the<br />

weathering <strong>an</strong>d alteration <strong>of</strong> previously deposited volc<strong>an</strong>ic ash <strong>an</strong>d colluvium, as well as other<br />

relatively s<strong>of</strong>t Pleistocene materials. Seepage <strong>of</strong> water from the surrounding soil into the<br />

mottled <strong>clays</strong> occurs during the rainy season, causing the latter soils to be water-logged for a<br />

period <strong>of</strong> 2 to 3 months, commonly with a water depth <strong>of</strong> up to 30 cm lying on the soil<br />

surface. The edges <strong>of</strong> the mottled <strong>clays</strong> are usually characterised by the formation <strong>of</strong> a laterite


46<br />

horizon, most probably a result <strong>of</strong> base accumulations from the seepage waters coming from<br />

the surrounding soils. The mottling character <strong>of</strong> these <strong>clays</strong> serves to distinguish <strong>an</strong>d separate<br />

them from the otherwise black to dark grey <strong>clays</strong> (black cotton soils) <strong>of</strong> the Athi <strong>an</strong>d Kapiti<br />

plains. The mottled <strong>clays</strong> are usually found uncultivated in their areas <strong>of</strong> occurrence. They<br />

therefore exhibit <strong>an</strong>d are recognised by a uniform light colour tone on aerial photos given by<br />

the grass vegetation cover.<br />

Deposits <strong>of</strong> pale brown or reddish ferruginous soil <strong>an</strong>d ferricrete (murram) are also found<br />

occurring, especially in the southern half <strong>of</strong> the present area. These materials are products <strong>of</strong><br />

weathering <strong>an</strong>d alteration <strong>of</strong> underlying lava <strong>an</strong>d tuff. They are generally hard, <strong>of</strong> sufficient<br />

thickness <strong>an</strong>d commonly quarried for use as road surfacing material. The south-western part<br />

<strong>of</strong> the study area, along L<strong>an</strong>gata road <strong>an</strong>d around Nairobi National Park, shows the Nairobi<br />

trachyte escarpment to be sufficiently weathered, giving rise to considerable deposits <strong>of</strong><br />

ferricrete. This material is extracted from a number <strong>of</strong> quarries located at the base <strong>of</strong> the<br />

scarp, <strong>an</strong>d utilised for surfacing earth roads <strong>an</strong>d tracks.<br />

5.2 Soils<br />

The Nairobi region, which includes the present study area, is covered by various thicknesses<br />

<strong>of</strong> soils that are products <strong>of</strong> weathering <strong>of</strong> mainly volc<strong>an</strong>ic rocks under relatively high<br />

temperature <strong>an</strong>d rainfall conditions. Good drainage conditions prevailing in the Kikuyu<br />

highl<strong>an</strong>ds to the west <strong>of</strong> Nairobi city as well as in the north-western section <strong>of</strong> the present<br />

study area, have facilitated the weathering processes to give rise to red soils <strong>of</strong> up to 15 m<br />

thickness. On the other h<strong>an</strong>d, poorer drainage conditions typifying the Athi <strong>an</strong>d Kapiti plains<br />

covering much <strong>of</strong> the present area to the south <strong>an</strong>d east <strong>of</strong> the Nairobi city, have favoured the<br />

development <strong>an</strong>d formation <strong>of</strong> black cotton soils.<br />

Early works in Nairobi <strong>an</strong>d the present study area classified the soils found by separating<br />

them into major world groups, based on the soil map in the Atlas <strong>of</strong> Kenya (Survey <strong>of</strong> Kenya,<br />

1991). Scott (1963), modified the classification to facilitate a survey carried out as <strong>an</strong> exercise<br />

in the interpretation <strong>of</strong> aerial photographs. In the present study, initial separation <strong>of</strong> soils was<br />

also largely based on actual observations made on aerial photographs. A further subdivision<br />

<strong>of</strong> the soils involved their classification according to climatic regions, topography <strong>an</strong>d/ or<br />

slope, <strong>an</strong>d drainage. Thus soils could be placed into one <strong>of</strong> three classes <strong>of</strong> well-drained soils,<br />

soils with slight seasonal impeded drainage or soils with impeded drainage. The well-drained<br />

soils could be further subdivided into soils <strong>of</strong> humid, sub-humid or semi-arid climatic regions<br />

<strong>an</strong>d/ or conditions. Humid regions have rainfall amounts <strong>of</strong> over 1000 mm per year; while<br />

sub-humid <strong>an</strong>d semi-arid regions are characterised by rainfall amounts <strong>of</strong> 760-1000 mm <strong>an</strong>d<br />

500-760 mm per year, respectively. The distribution <strong>of</strong> major soil types found in Nairobi <strong>an</strong>d<br />

the present study area is presented in Figure 5.1. The soil map serves to give a good general<br />

picture <strong>of</strong> the study area <strong>an</strong>d its surroundings. It also serves as a guide to where more detailed<br />

soil studies could be undertaken, as may be required. A geotechnical soil map (Appendix E)<br />

has also been prepared; <strong>an</strong>d apart from showing the distribution, extent, boundaries <strong>an</strong>d depth<br />

variation <strong>of</strong> red soils <strong>an</strong>d black <strong>clays</strong>, it also presents a summary <strong>of</strong> the index <strong>an</strong>d <strong>engineering</strong><br />

properties <strong>of</strong> the soils (details discussed in a later chapter).


48<br />

5.2.1 Red friable <strong>clays</strong><br />

The red soils are found overlying the trachytic rocks in the higher altitude sections <strong>of</strong> the<br />

study area, located mainly to the north-west. This part <strong>of</strong> the study area is sub-humid <strong>an</strong>d<br />

characterised by a high rainfall <strong>of</strong> 760 – 1000 mm per year (Kenya Meteorological<br />

Department, 2001). The soils are well-drained <strong>an</strong>d exhibit a dark reddish-brown (Sample No.<br />

Rd-30cm) <strong>an</strong>d humic (up to 2,5% carbon) topsoil overlying a red subsoil (Sample Nos. Rd-<br />

100cm; Rd-200cm; Rd-400cm) <strong>of</strong> weak sub<strong>an</strong>gular blocky friable clay. Stone lines are<br />

sometimes found associated, occurring just above the metamorphic rocks.<br />

The soils are most probably polygenetic in that, the humid conditions must have favoured<br />

their development by weathering <strong>an</strong>d alteration from the volc<strong>an</strong>ic rocks, tuff <strong>an</strong>d ash found.<br />

Later episodes <strong>of</strong> violent volc<strong>an</strong>icity in the Rift region to the west must have contributed<br />

additional showers <strong>of</strong> volc<strong>an</strong>ic ash that covered already developed red soils <strong>of</strong> the underlying<br />

l<strong>an</strong>d surface (Scott, 1963). Current field observations show the red soils as having pr<strong>of</strong>iles<br />

consisting <strong>of</strong> burried humic horizons; <strong>an</strong>d this serves to indicate that sufficient time must have<br />

elapsed between subsequent ash showers, thereby permitting the development <strong>an</strong>d formation<br />

<strong>of</strong> a new soil from the previously deposited ash.<br />

The red soils extend further westwards <strong>an</strong>d north-westwards, beyond the present area <strong>an</strong>d<br />

across the Kikuyu highl<strong>an</strong>ds, to cover <strong>an</strong>d include Kiambu township. The soils occupy<br />

principally the tops /summits as well as upper <strong>an</strong>d middle slopes <strong>of</strong> hills <strong>an</strong>d ridges found, <strong>an</strong>d<br />

this serves to indicate the irregular character <strong>of</strong> the l<strong>an</strong>d surface over which the volc<strong>an</strong>ic ash<br />

fell. On the lower slopes <strong>of</strong> the ridges, the topography becomes generally flatter, with<br />

accomp<strong>an</strong>ying poorer drainage conditions, resulting in the formation <strong>of</strong> shallow yellow-brown<br />

to yellow-red friable <strong>clays</strong> overlying a laterite horizon or rock. The tr<strong>an</strong>sition zone between<br />

the red friable <strong>clays</strong> <strong>an</strong>d the shallow yellow-brown to yellow-red friable <strong>clays</strong> is characterised<br />

by narrow b<strong>an</strong>ds <strong>of</strong> <strong>an</strong> intermediate red friable clay with iron concretions, merging into<br />

massive laterite.<br />

Commonly found associated with the red friable <strong>clays</strong> are intercalations <strong>of</strong> poorly drained<br />

dark greyish-brown mottled <strong>clays</strong> observable in parts <strong>of</strong> the Nairobi municipality; <strong>an</strong>d which<br />

also occur in depressions between ridges, as well as the generally flat l<strong>an</strong>d surface overlying<br />

the trachytes to the north <strong>an</strong>d north-west <strong>of</strong> Nairobi city. The mottled <strong>clays</strong> are partly a result<br />

<strong>of</strong> the development <strong>of</strong> subsidiary drainage in the lowlying l<strong>an</strong>d <strong>an</strong>d/ or depressions between<br />

the undulations <strong>of</strong> ridges.<br />

Further north-westwards, beyond the present area <strong>of</strong> study, the red soils give way to strong<br />

brown to yellow-red friable <strong>clays</strong> <strong>an</strong>d dark red friable <strong>clays</strong> in a humid region covering the<br />

high ground fl<strong>an</strong>king the Rift Valley, <strong>an</strong>d where a rainfall amount <strong>of</strong> over 1000 mm per year<br />

is characteristic (Kenya Meteorological Department, 2001). These latter types <strong>of</strong> soil also<br />

overly the trachytic rocks found, <strong>an</strong>d have been derived from the volc<strong>an</strong>ic rocks, tuff <strong>an</strong>d ash<br />

through processes <strong>of</strong> weathering <strong>an</strong>d alteration. The soils are well-drained <strong>an</strong>d also<br />

characterised by a dark reddish-brown high humic (3-7% carbon) topsoil overlying a dark red<br />

subsoil <strong>of</strong> sub<strong>an</strong>gular blocky friable clay with clay skins. The red friable <strong>clays</strong> are<br />

comparatively less humic, have a lower total exch<strong>an</strong>geable base content <strong>an</strong>d are less saturated<br />

th<strong>an</strong> the dark red friable <strong>clays</strong>. They are also characterised by a much weaker structure <strong>an</strong>d<br />

less marked development <strong>of</strong> clay skins. The brighter red colour <strong>an</strong>d weaker structure <strong>of</strong> these<br />

soils may be partly a result <strong>of</strong> the lower humus status.


49<br />

The red friable <strong>clays</strong> on the whole are relatively deeper th<strong>an</strong> the black <strong>clays</strong>, mottled <strong>clays</strong> <strong>an</strong>d<br />

shallow yellow-brown to yellow-red friable <strong>clays</strong>. This may possibly be attributed to the<br />

higher rainfall received in areas <strong>of</strong> red soils as well as better drainage conditions <strong>of</strong> these soils<br />

both <strong>of</strong> which favour a relatively faster weathering <strong>of</strong> the underlying rock. However,<br />

development <strong>of</strong> the red soils on the concave slopes <strong>of</strong> the ridges is limited by the formation <strong>of</strong><br />

a laterite cap over parts <strong>of</strong> the rock surface; <strong>an</strong>d this would protect the rock from further<br />

weathering. The formation <strong>of</strong> laterite is favoured by the generally concave shape <strong>of</strong> the slopes<br />

<strong>of</strong> the ridges that results in ch<strong>an</strong>ging drainage conditions downslope.<br />

Table 5.1. Pr<strong>of</strong>ile description <strong>of</strong> a red friable clay.<br />

DEPTH<br />

(M)<br />

SAMPLE<br />

NO.<br />

DESCRIPTION<br />

0,0-0,30 Rd1-30cm Dark reddish brown, slightly<br />

humic, weak crumbly friable<br />

clay<br />

0,30-1,0 Rd1-100cm Dark red to red weak<br />

sub<strong>an</strong>gular blocky friable clay<br />

1,0-2,0 Rd1-200cm Red weak sub<strong>an</strong>gular blocky<br />

friable clay with slight clay skin<br />

development<br />

2,0-4,0 Rd1-400cm Red weak sub<strong>an</strong>gular blocky<br />

friable clay with slight clay skin<br />

development<br />

SILT<br />

(%)<br />

CLAY<br />

(%)<br />

C (%)<br />

73 18 2,47<br />

74 20 0,73<br />

69 24 0,83<br />

72 20 0,45<br />

The boundary lines between the red <strong>clays</strong> <strong>an</strong>d other clay soils mentioned above were first<br />

fixed along roads <strong>an</strong>d/ or pl<strong>an</strong>ned traverses by ground inspection; <strong>an</strong>d then extended <strong>an</strong>d<br />

joined by interpolation. On the aerial photographs, the red <strong>clays</strong> show up as darker tones<br />

relative to the other adjacent soil types mentioned above. A typical red friable clay from the<br />

present study area is described in Table 5.1. The amount <strong>of</strong> total carbon was found to decrease<br />

with depth, <strong>an</strong>d this is most likely also indicative <strong>of</strong> decreasing org<strong>an</strong>ic matter (humus)<br />

content <strong>of</strong> the soils with depth.<br />

Table 5.2. Soluble base content in soils <strong>of</strong> the study area.<br />

Type <strong>of</strong> soil Depth<br />

(m)<br />

Sample<br />

No.<br />

MgO<br />

(%)<br />

CaO<br />

(%)<br />

MnO<br />

(%)<br />

Na2O<br />

(%)<br />

K2O<br />

(%)<br />

Total<br />

(%)<br />

Red friable 0,0-0,30 Rd1-30cm 0,12 0,26 0,96 - 0,96 2,30<br />

<strong>clays</strong> 0,30-1,0 Rd1-100cm 0,097 0,24 0,53 - 0,66 1,53<br />

1,0-2,0 Rd1-200cm 0,14 0,16 0,61 - 0,68 1,59<br />

2,0-4,0 Rd1-400cm 0,13 0,26 0,52 - 0,60 1,51<br />

Shallow yellow 0,0-0,30 SC17-30cm 1,17 1,28 0,529 0,74 1,33 5,05<br />

- brown <strong>clays</strong> 0,30-0,50 SC17-50cm 1,28 1,47 0,488 0,82 1,41 5,47<br />

on laterite >0,50 - - - - - - -<br />

0,0-0,30 SA2-30cm 1,07 1,25 0,298 0,63 1,30 4,55<br />

Black <strong>clays</strong> 0,30-0,70 SA2-70cm 1,25 1,28 0,312 1,06 1,78 5,68<br />

0,70-1,05 SA2-105cm 1,22 4,29 0,363 1,04 1,73 8,64<br />

Black <strong>clays</strong> 0,0-0,30 SB1-30cm 1,02 1,26 0,44 0,70 1,35 4,77<br />

0,30-0,50 SB1-50cm 1,16 1,54 0,455 0,71 1,27 5,14<br />

0,50-0,70 SB1-70cm 1,23 2,01 0,439 0,76 1,34 5,78


50<br />

The variation in the content <strong>of</strong> total soluble bases in red soils is presented in Table 5.2<br />

(sample Rd1), alongside that for black <strong>clays</strong> for comparison purposes. The amount <strong>of</strong> total<br />

soluble base was also observed to decrease with increasing depth <strong>of</strong> the red soils. This could<br />

be explained by the good drainage conditions <strong>of</strong> the soils which serve to facilitate removal <strong>of</strong><br />

the basic subst<strong>an</strong>ces in solution during rainfall.<br />

Further eastwards across the study area, the red soils pass into a composite soil group<br />

consisting <strong>of</strong> shallow yellow-brown friable <strong>clays</strong> overlying a laterite horizon; <strong>an</strong>d shallow<br />

yellow-red friable <strong>clays</strong> on rock. This is because <strong>of</strong> the difficulty in distinguishing between<br />

<strong>an</strong>d separating these two soil units on aerial photographs. However, the two soil components<br />

are genetically different. These soils commonly exhibit a brown (Sample No. SC17-30cm)<br />

relatively low humic (up to 1,5% carbon) topsoil overlying a yellow-brown (Sample No.<br />

SC17-50cm) sub<strong>an</strong>gular blocky friable clay with iron concretions passing downwards into<br />

massive laterite (ferricrete). In other places, the soils consist <strong>of</strong> a brown low humic (0,5-1,5%<br />

carbon) topsoil underlain by a yellow-red sub<strong>an</strong>gular blocky friable clay passing into rock.<br />

The soils are usually characterised by a slight seasonal impeded drainage. They also support a<br />

scrub grass vegetation which appears as a light tone on aerial photographs. The shallow soils<br />

over rock are generally redder th<strong>an</strong> their lateritic counterparts. They are found occurring<br />

mainly on the flat l<strong>an</strong>ds adjoining <strong>an</strong>d/ or at the edges <strong>of</strong> the black <strong>clays</strong> in the plains. They<br />

are apparently youthful soils which have developed most probably after the removal <strong>of</strong> the<br />

black <strong>clays</strong> by erosive processes. The shallow soils over laterite occur mainly on the lower<br />

slopes <strong>of</strong> ridges. According to Scott (1963), they are most probably a result <strong>of</strong> seepage water<br />

from higher ground whose downslope flow is checked by the ch<strong>an</strong>ge <strong>of</strong> slope <strong>an</strong>d poorer<br />

drainage conditions at the foot <strong>of</strong> the slope, causing the deposition <strong>of</strong> iron <strong>an</strong>d aluminium<br />

compounds from the seepage waters to form a laterite sheet. The development <strong>of</strong> the laterite<br />

sheet also serves to increase the drainage impedence , thereby favouring further development<br />

<strong>an</strong>d formation <strong>of</strong> the shallow lateritic soils which extend upslope.<br />

Table 5.3. Pr<strong>of</strong>ile description <strong>of</strong> a shallow yellow-brown clay overlying laterite.<br />

Depth (m) Sample No. Description Silt Clay C (%)<br />

(%) (%)<br />

0,0-0,30 SC17-30cm Brown slightly humic,<br />

crumbly friable clay with<br />

occasional rounded murram<br />

concretion<br />

59 29 1,54<br />

0,30-0,50 SC17-50cm Yellow brown sub<strong>an</strong>gular<br />

blocky friable clay with<br />

occasional rounded murram<br />

concretion<br />

>0,50 - Mainly murram concretion<br />

passing into massive murram<br />

59 27 1,29<br />

- - -<br />

A typical yellow-brown friable clay overlying laterite in the current study area is described in<br />

Table 5.3. The amount <strong>of</strong> total carbon is observed to decrease with increasing depth <strong>of</strong> soils,<br />

<strong>an</strong>d this could as well serve to reflect possible decrease <strong>of</strong> org<strong>an</strong>ic matter content <strong>of</strong> the soils<br />

with depth. However, the soils exhibit a slight increase in the amount <strong>of</strong> total soluble base<br />

with increased depths (Table 5.2, Sample SC17), <strong>an</strong>d this is most probably due to the slightly<br />

impeded drainage character <strong>of</strong> the soils which serves to limit <strong>an</strong>d/ or inhibit removal <strong>of</strong> the<br />

basic subst<strong>an</strong>ces in solution.


51<br />

Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> the red soils performed in the current study (Table 5.4) show<br />

them as lacking <strong>an</strong>d deficient in magnesia (MgO: 0,097-0,14 %). Chemical <strong>an</strong>alyses <strong>of</strong> the<br />

rocks from which these soils are derived also indicate their relatively very low magnesia<br />

content (MgO: 0,30-0,89 %), compared with the basic lavas <strong>an</strong>d other volc<strong>an</strong>ic rocks <strong>of</strong> the<br />

Nairobi region (Table 3.4). Farmers in the c<strong>of</strong>fee-growing estates found in the areas <strong>of</strong> red<br />

soils tackle this problem through frequent artificial application <strong>of</strong> magnesite-rich fertilisers.<br />

The red soils generally exhibit lower contents <strong>of</strong> SiO2, MgO, CaO, Na2O <strong>an</strong>d K2O th<strong>an</strong> the<br />

volc<strong>an</strong>ic materials from which they are derived (Tables 3.3 <strong>an</strong>d 5.4). This could also be a<br />

result <strong>of</strong> the good drainage conditions <strong>of</strong> the soils as well as the high rainfall which favoured<br />

leaching <strong>an</strong>d removal <strong>of</strong> the soluble bases <strong>an</strong>d silica, thereby leaving the soils relatively<br />

enriched in Fe2O3 <strong>an</strong>d Al2O3 minerals.<br />

Table 5.4. Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> red soils obtained in this study.<br />

% content <strong>of</strong>: Rd1-30cm Rd1-100cm Rd1-200cm Rd1-400cm<br />

SiO2 47,50 47,10 47,20 47,6<br />

Al2O3 32,20 33,40 33,20 33,30<br />

Fe2O3 15,30 15,50 15,50 15,20<br />

FeO - - - -<br />

BaO 0,062 0,061 0,071 0,064<br />

MgO 0,12 0,097 0,14 0,13<br />

CaO 0,26 0,24 0,16 0,26<br />

Na2O ‹ ‹ ‹ ‹<br />

K2O 0,96 0,66 0,68 0,60<br />

P2O5 0,19 0,086 0,096 0,062<br />

ZrO2 0,37 0,37 0,37 0,36<br />

TiO2 1,57 1,51 1,51 1,46<br />

MnO 0,96 0,53 0,61 0,52<br />

Loss on ignition - - - -<br />

Total 99,49 99,55 99,54 99,56<br />

Rd1-30cm, Rd1-100cm, Rd1-200cm, Rd1-400cm. Samples <strong>of</strong> red soils collected from Arboretum (Nairobi), at<br />

depths <strong>of</strong> 30cm, 100cm, 200cm <strong>an</strong>d 400cm, respectively.<br />

‹: Concentration less th<strong>an</strong> 0,04 % (400 ppm).<br />

Similar results <strong>of</strong> previous chemical <strong>an</strong>alyses <strong>of</strong> a number <strong>of</strong> <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> Nairobi<br />

area are also given in Table 5.5, for comparison purposes. A relation <strong>of</strong> the results <strong>of</strong> Tables<br />

5.4 <strong>an</strong>d 5.5 reveals a depletion in the amounts <strong>of</strong> soluble bases (MgO, CaO, NaO) to have<br />

taken place with time, most probably due to effects <strong>of</strong> leaching in the red soils.<br />

Correspondingly, the proportionate amounts <strong>of</strong> Fe2O3 (free iron oxide, haematite, goethite)<br />

<strong>an</strong>d Al2O3 (clay minerals) have increased with time.<br />

The more easily leached components from areas <strong>of</strong> red soils persist or accumulate on lower<br />

ground <strong>of</strong> depressions between ridges <strong>an</strong>d/ or the flat plains, where they contribute to the<br />

formation <strong>of</strong> black soils.


52<br />

Table 5.5. Earlier results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> <strong>clays</strong> <strong>an</strong>d red soil.<br />

% content <strong>of</strong>: 1 2 3 4<br />

SiO2 40,65 34,16 38,69 35,60<br />

Al2O3 20,98 25,55 32,74 30,75<br />

Fe2O3 11,48 14,15 13,56 15,55<br />

FeO - 0,90 - -<br />

MgO 1,50 0,37 0,26 -<br />

CaO 2,00 1,09 Tr -<br />

Na2O - 0,09 Nd -<br />

K2O - 0,34 Nd -<br />

H2O+ 9,62 9,24 - -<br />

H2O- 12,51 11,10 - -<br />

TiO2 - 3,12 1,07 1,40<br />

MnO - 0,38 0,19 -<br />

Loss on ignition - - 13,14 15,03<br />

Total 98,74 100,49 99,65 98,33<br />

1. Clay. Fairview Estate, Kiambu. .<br />

2. Red clay. Escarpment, Kiambu (Maufe, 1908)<br />

3. Red soil, Kiambu (EAIRB, 1953. Technical Pamphlet 61)<br />

4. Clay. Sasamua (Terzaghi, 1958).<br />

According to Rösler (1980), weathering <strong>an</strong>d alteration <strong>of</strong> feldspar-rich volc<strong>an</strong>ic materials<br />

under humid conditions <strong>of</strong> rainfall <strong>an</strong>d in the presence <strong>of</strong> carbon dioxide usually results in the<br />

conversion <strong>of</strong> aluminiumsilicates contained, <strong>an</strong>d formation <strong>of</strong> kaolinite-rich soils. The red<br />

soils in the present area must have therefore mainly formed by weathering <strong>an</strong>d alteration <strong>of</strong><br />

the underlying Nairobi trachytes. This is evidenced by chemical/ mineralogical studies <strong>of</strong> the<br />

soils which show them as being predomin<strong>an</strong>tly kaolinite (80-81%) <strong>an</strong>d iron oxide (haematite;<br />

15-16%) in their composition (see next chapter). The mineral components rich in alkali (Na,<br />

K) <strong>an</strong>d alkali-earth metals (Mg, Ca), as well as silica (SiO2) were therefore removed in<br />

solution, leaving behind the altered volc<strong>an</strong>ic rocks rich in quartz, aluminium <strong>an</strong>d other stable<br />

minerals in the form <strong>of</strong> kaolinite-rich red soils. These findings are further supported by<br />

previous studies made by staff <strong>of</strong> the Ministry <strong>of</strong> Agriculture on the red clay soils <strong>of</strong> Nairobi,<br />

which revealed that leaching <strong>of</strong> the soils had the effect <strong>of</strong> removing the soluble bases <strong>an</strong>d<br />

silica, leaving the soils rich in iron oxide <strong>an</strong>d/or haematite; as well as aluminium in the form<br />

<strong>of</strong> clay minerals, metahalloysite <strong>an</strong>d hydrated halloysite (Dumbleton, 1967; Sherwood, 1967).<br />

The free iron oxide has the effect <strong>of</strong> cementing the clay particles together. This is a possible<br />

expl<strong>an</strong>ation for the <strong>an</strong>omalous results obtained in the present study by subjecting the red soils<br />

to British St<strong>an</strong>dard soil classification tests (Table 5.6).<br />

Table 5.6. Some results <strong>of</strong> soil classification tests performed on red soils in the present study.<br />

LOCATION SAMPLE<br />

NO.<br />

CLAY<br />

CONTENT<br />

(BS 1377)<br />

(%)<br />

LIQUID<br />

LIMIT<br />

(%)<br />

PLASTIC<br />

LIMIT<br />

(%)<br />

PLASTICITY<br />

INDEX<br />

(%)<br />

Arboretum Rd1-30cm 18 49 31 18<br />

(Nairobi)<br />

Arboretum Rd1-100cm 20 51 31 20<br />

Arboretum Rd1-200cm 24 50 30 20<br />

Arboretum Rd1-400cm 20 48 30 18


53<br />

Results <strong>of</strong> previous soil classification tests carried out on a number <strong>of</strong> <strong>clays</strong> <strong>an</strong>d red soils in<br />

the neighbourhood <strong>of</strong> the present study area are summarised in Table 5.7, for comparison<br />

purposes.<br />

It c<strong>an</strong> be observed from the results <strong>of</strong> Tables 5.6 <strong>an</strong>d 5.7 that the cementing effect <strong>of</strong> clay<br />

particles by free iron oxide must have progressively reduced the proportion <strong>of</strong> free active clay<br />

particles in the red soils through <strong>geological</strong> time. As a result, values <strong>of</strong> the Atterberg limits<br />

(plastic <strong>an</strong>d liquid limits, plasticity indices) <strong>of</strong> the soils have also decreased with time.<br />

Table 5.7. Earlier results <strong>of</strong> soil classification tests on red <strong>clays</strong> from the Nairobi area (After<br />

Sherwood, 1967).<br />

LOCATION<br />

SOIL<br />

NO.<br />

CLAY<br />

CONTENT<br />

(BS1377)<br />

(%)<br />

LIQUID<br />

LIMIT<br />

(%)<br />

PLASTIC<br />

LIMIT<br />

(%)<br />

PLASTICITY<br />

INDEX<br />

(%)<br />

Kabete 816 82 76 39 37<br />

Limuru 830 83 87 44 43<br />

Nairobi-Limuru<br />

road<br />

639 88 87 48 39<br />

5.2.2 Black to dark grey <strong>clays</strong> (black cotton soils)<br />

The black to dark grey <strong>clays</strong> cover much <strong>of</strong> the present study area that forms a part <strong>of</strong> the Athi<br />

<strong>an</strong>d Kapiti plains. The soils are characterised by impeded drainage, <strong>an</strong>d are underlain mainly<br />

by the Nairobi <strong>an</strong>d Kapiti phonolites which form a practically impermeable strata, favouring<br />

further development <strong>an</strong>d occurrence <strong>of</strong> the ill-drained soils. Contributions to the formation <strong>of</strong><br />

the black <strong>clays</strong> also comes from eroded kaolinite-bearing materials as well as leached<br />

components <strong>of</strong> soluble bases <strong>an</strong>d silica derived from areas <strong>of</strong> red soils occurring on the higher<br />

grounds forming the Kikuyu highl<strong>an</strong>ds to the west <strong>an</strong>d north-west <strong>of</strong> the present area. This is<br />

evidenced by chemical/ mineralogical results from current studies, which show the black<br />

<strong>clays</strong> as containing relatively elevated amounts <strong>of</strong> MgO, CaO, Na2O, K2O <strong>an</strong>d SiO2 (Table<br />

5.8), as well as traces <strong>of</strong> finely dispersed kaolinite (Table 5.9).<br />

The black to dark grey <strong>clays</strong> are also known as black cotton soils which are characterised by<br />

both calcareous <strong>an</strong>d non-calcareous vari<strong>an</strong>ts. The calcareous types occur on a much lesser<br />

extent in the south-eastern sections <strong>of</strong> the study area, while the non-calcareous varieties cover<br />

the remaining portion <strong>of</strong> the plains in the present study area.<br />

Commonly found associated are local swampy environments <strong>an</strong>d alluvium, which occur as<br />

isolated patches within the black cotton soils. A limited intercalated occurrence <strong>of</strong> the dark<br />

greyish-brown mottled <strong>clays</strong> is also recognised, while shallow stony soils with rock outcrops<br />

are concentrated along the generally easterly/ south-easterly trending stream valleys found.


54<br />

Table 5.8. Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> black <strong>clays</strong> obtained in this study.<br />

% SA2- SA2- SA41- SB1- SB1- SB41- SB41- SC17-<br />

content<br />

<strong>of</strong>:<br />

70cm 105cm 50cm 50cm 70cm 30cm 50cm 50cm<br />

SC41-<br />

30cm<br />

SC41-<br />

50cm<br />

SiO2 52,50 51,40 58,35 52,79 53,44 52,64 53,21 56,38 52,06 53,71<br />

Al2O3 14,71 14,93 13,37 12,74 12,84 12,14 12,20 11,87 12,06 11,77<br />

Fe2O3 8,66 8,85 6,98 6,90 6,96 6,67 6,75 6,20 6,49 6,10<br />

FeO - - - - - - - - - -<br />

BaO 0,052 0,052 0,031 0,062 0,062 0,064 0,062 0,085 0,123 0,104<br />

MgO 1,25 1,22 1,13 1,16 1,23 1,43 1,42 1,28 1,34 1,31<br />

CaO 1,28 4,29 1,28 1,54 2,01 2,61 2,32 1,47 1,64 1,48<br />

Na2O 1,06 1,04 0,69 0,71 0,76 0,50 0,56 0,82 0,90 0,89<br />

K2O 1,78 1,73 1,25 1,27 1,34 1,02 1,04 1,41 1,32 1,31<br />

P2O5 0,026 0,024 0,037 0,032 0,038 0,038 0,038 0,047 0,036 0,04<br />

ZrO2 0,092 0,091 0,085 0,091 0,092 0,078 0,078 0,072 0,073 0,072<br />

TiO2 0,96 0,95 0,90 0,76 0,75 0,77 0,77 0,75 0,79 0,77<br />

MnO 0,312 0,363 0,249 0,455 0,439 0,497 0,505 0,488 0,469 0,511<br />

SO3 - - - - - - - 0,011 0,028 0,016<br />

Cl 0,065 0,039 - - - - - - 0,015 0,015<br />

F 0,262 0,362 0,247 0,294 0,325 0,281 0,267 0,311 0,300 0,293<br />

Loss on - - - - - - - - - -<br />

ignition<br />

Total 83,01 85,34 84,60 78,80 80,29 78,74 79,22 81,19 77,64 78,39<br />

SA2-70cm, SA2-105cm, SA41-50cm, SB1-50cm, SB1-70cm, SB41-30cm, SB41-50cm, SC17-50cm, SC41-<br />

30cm <strong>an</strong>d SC41-50cm are black clay samples.<br />

The black to dark grey <strong>clays</strong> are characterised by a number <strong>of</strong> distinctive properties which<br />

serve to point to their possible genetic origin, as discussed below.<br />

Results <strong>of</strong> chemical <strong>an</strong>alyses performed on these soils in this study (Table 5.8) show the<br />

common exch<strong>an</strong>geable bases <strong>of</strong> Ca, Mg, K <strong>an</strong>d Na to be all <strong>of</strong> the same order irrespective <strong>of</strong><br />

the type <strong>of</strong> rock underlying the soil. This serves to show that the black <strong>clays</strong> are most<br />

probably not derived from the underlying rocks by weathering. This is further supported by<br />

results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> the soils, which generally do not agree with <strong>an</strong>d/ or are not <strong>of</strong><br />

the same order as those for the main underlying rocks <strong>of</strong> Nairobi <strong>an</strong>d Kapiti phonolites<br />

(Tables 3.1, 3.2 <strong>an</strong>d 3.3).<br />

The black <strong>clays</strong> exhibit a generally uniform depth, varying from 0,90 to 1,50m. Depths <strong>of</strong> up<br />

to 1,80 to 2,10m are also registered in the swampy areas <strong>of</strong> Madaraka West <strong>an</strong>d environs. The<br />

<strong>clays</strong> generally show pronounced cracking in the form <strong>of</strong> strong shrinkage cracks on drying<br />

from a wet condition.<br />

The black <strong>clays</strong> were also observed harbouring occasional s<strong>an</strong>dier vari<strong>an</strong>ts in some places.<br />

Results <strong>of</strong> X-ray diffraction <strong>an</strong>alyses performed in this study as well as earlier results <strong>of</strong><br />

chemical/ mineralogical studies by Stephen, Bellis <strong>an</strong>d Muir (1956) show the black <strong>clays</strong> as<br />

containing a number <strong>of</strong> heavy minerals usually characteristic <strong>of</strong> metamorphic rocks <strong>of</strong> the<br />

gneissic type. The heavy minerals include sillim<strong>an</strong>ite, ky<strong>an</strong>ite, garnet <strong>an</strong>d staurolite; <strong>an</strong>d were<br />

traced in soil samples collected from Embakasi area <strong>an</strong>d its environs which are underlain by<br />

Nairobi phonolite. The presence <strong>of</strong> the s<strong>an</strong>dier vari<strong>an</strong>ts <strong>an</strong>d heavy minerals serves to point to<br />

the possibility <strong>of</strong> portions <strong>of</strong> materials forming the black <strong>clays</strong> as having originated from


55<br />

neighbouring metamorphic areas <strong>an</strong>d deposited by rivers <strong>an</strong>d/ or surface run-<strong>of</strong>f into the<br />

present area. Alteration <strong>of</strong> kaolinitic red soils deficient in alkali <strong>an</strong>d alkali earth metals (Na,<br />

K, Mg, Ca) could have also caused the formation <strong>of</strong> the respective aluminiumsilicates <strong>of</strong> the<br />

heavy minerals. The altered soils were then eroded, tr<strong>an</strong>sported <strong>an</strong>d deposited in the lower<br />

areas <strong>of</strong> black <strong>clays</strong>.<br />

The black <strong>clays</strong> generally exhibit sharp to fairly sharp contacts with the underlying volc<strong>an</strong>ic<br />

rocks. Cases <strong>of</strong> patches <strong>of</strong> black <strong>clays</strong> overlying red soils were observed in the northern <strong>an</strong>d<br />

north-western sections <strong>of</strong> the study area. In other parts <strong>of</strong> the study area, laterite was<br />

commonly found underlying the black <strong>clays</strong>. In addition, areas with black <strong>clays</strong> having been<br />

removed by erosion show the underlying rock weathering through red. The above four<br />

observations contradict the suggestion that the black <strong>clays</strong> could have been derived from the<br />

underlying volc<strong>an</strong>ic rocks by weathering.<br />

The neighbouring areas further eastwards are underlain by metamorphic rocks, <strong>an</strong>d commonly<br />

exhibit a line <strong>of</strong> quartz stones between the black soils <strong>an</strong>d rock (Scott, 1963; Saggerson,<br />

1991). However, no quartz veins have been observed within the black <strong>clays</strong>; <strong>an</strong>d this is<br />

contrary to what would be expected if the soils had actually developed in situ, suggesting<br />

once again that the black <strong>clays</strong> may have not derived from the weathering <strong>of</strong> the underlying<br />

rocks. In the present study area underlain by volc<strong>an</strong>ic rocks, pieces <strong>of</strong> quartz stones are found<br />

r<strong>an</strong>domly imbedded in the black <strong>clays</strong>. This could point to the possibility <strong>of</strong> the quartz stones<br />

<strong>an</strong>d other materials as having been tr<strong>an</strong>sported from the surrounding areas into the present<br />

area by erosive <strong>an</strong>d depositional processes.<br />

The general morphology <strong>of</strong> the area <strong>of</strong> occurrence <strong>of</strong> black <strong>clays</strong> in the present study area <strong>an</strong>d<br />

neighbouring areas further eastwards is roughly basin-like. The eastern <strong>an</strong>d western margins<br />

<strong>of</strong> this basin tend to follow the 5200 feet contour, while the northern margin is at <strong>an</strong> altitude<br />

<strong>of</strong> about 4800 feet. The altitude <strong>of</strong> the southern boundary <strong>of</strong> the soils is up to 6200 feet in the<br />

vicinity <strong>of</strong> Ngong hills. The soils are found occurring on both flat sites <strong>an</strong>d appreciable slopes.<br />

However, isolated patches <strong>of</strong> black <strong>clays</strong> occurring on high grounds are also found; <strong>an</strong>d are<br />

possibly part <strong>of</strong> the same earlier <strong>an</strong>d larger formation, which later underwent erosion leaving<br />

behind only relics <strong>of</strong> the soils.<br />

The properties <strong>an</strong>d characteristics <strong>of</strong> black <strong>clays</strong> discussed above serve to conclusively<br />

suggest that these soils are not wholly developed in situ. A more plausible expl<strong>an</strong>ation as to<br />

the genetic origin <strong>of</strong> the black <strong>clays</strong> would be based on the exist<strong>an</strong>ce <strong>of</strong> a large basin in their<br />

present areas <strong>of</strong> occurr<strong>an</strong>ce, prior to the formation <strong>of</strong> the soils. The basin was bounded by the<br />

Machakos hills to the east, the Ngong hills to the south <strong>an</strong>d the high ground <strong>of</strong> the eastern<br />

fl<strong>an</strong>ks <strong>of</strong> the Rift Valley to the west. The northern margin <strong>of</strong> the basin was most probably<br />

associated with the Ith<strong>an</strong>ga-Kakuzi hills. The lowest point <strong>of</strong> the basin was most probably<br />

around Athi River Township. The basin was filled with water during a wet cycle, forming a<br />

lake. Materials <strong>of</strong> colluvium <strong>an</strong>d alluvium from the surrounding hills were later deposited<br />

into the lake. Portions <strong>of</strong> the basin within <strong>an</strong>d/ or in the vicinity <strong>of</strong> metamorphic rock areas<br />

were therefore characterised by more s<strong>an</strong>dier vari<strong>an</strong>ts <strong>of</strong> black <strong>clays</strong> th<strong>an</strong> areas <strong>of</strong> the basin<br />

within volc<strong>an</strong>ic rock areas <strong>an</strong>d/ or next to volc<strong>an</strong>ic hill masses. This is also a possible<br />

expl<strong>an</strong>ation for the presence <strong>of</strong> s<strong>an</strong>dier vari<strong>an</strong>ts <strong>of</strong> black <strong>clays</strong> in the current study area, which<br />

is underlain by volc<strong>an</strong>ic rocks. In addition, there is a high possibility <strong>of</strong> volc<strong>an</strong>ic ash having<br />

also been deposited into the lake during this period; <strong>an</strong>d this could account for the high<br />

presence <strong>of</strong> smectites in the black <strong>clays</strong>, as detected during mineralogical (X-ray diffraction)<br />

<strong>an</strong>alyses performed on the soils in this study (Table 5.9).


56<br />

Table 5.9. Clay mineral composition <strong>of</strong> black <strong>clays</strong>.<br />

Sample No. Smectites<br />

(montmorillonite)<br />

(%)<br />

Illite<br />

(%)<br />

Kaolinite<br />

(%)<br />

Quartz<br />

(%)<br />

K-feldspars<br />

(s<strong>an</strong>idine,<br />

orthoclase)<br />

SA2-70cm >90 - 90 Trace 95 - 95 - 95 - 95 Trace 95 - 95 - >5 4 trace<br />

SC33-30cm >95 - 95 Trace


57<br />

Table 5.10. Pr<strong>of</strong>ile description <strong>of</strong> black clay.<br />

Depth (m) Sample No. Description Silt (%) Clay (%) C (%)<br />

0,0-0,30 SA2-30cm Very dark grey humic, <strong>an</strong>gular 64 26 0,98<br />

blocky plastic clay, cracks<br />

when dry<br />

0,30-0,7 SA2-70cm Dark grey <strong>an</strong>gular blocky 62 26 0,50<br />

plastic clay<br />

0,70-1,05 SA2-105cm Grey structureless plastic clay 61 24 0,96<br />

with occasional CaCO3<br />

concretions<br />

>1,05 - Murram <strong>an</strong>d/ or weathered tuff - -<br />

passing into phonolite rock<br />

0,0-0,30 SB1-30cm Very dark grey humic, <strong>an</strong>gular<br />

blocky plastic clay, cracks<br />

when dry<br />

0,30-0,6 SB1-50cm Dark grey <strong>an</strong>gular blocky<br />

plastic clay<br />

0,60-0,90 SB1-70cm Grey plastic clay with some<br />

CaCO3 concretions<br />

>0,90 - Weathered tuff passing into<br />

phonolite rock<br />

60 30 -<br />

65 26 1,13<br />

66 25 1,19<br />

- - -<br />

The black to dark grey <strong>clays</strong> appeared as dark tones on the aerial photos, with lighter toned<br />

mounds found scattered within being very prominent.<br />

The numbering <strong>of</strong> soils in this study follows the legend <strong>of</strong> maps to this thesis work.


58<br />

Chapter 6<br />

Chemical <strong>an</strong>d mineralogical <strong>an</strong>alyses<br />

6.1 Introduction<br />

A number <strong>of</strong> soil samples previously collected from the field were subjected to laboratory<br />

chemical <strong>an</strong>d mineralogical <strong>an</strong>alyses <strong>an</strong>d/ or studies using the X-ray fluorescence (XRF), X-<br />

ray diffraction (XRD) <strong>an</strong>d sc<strong>an</strong>ning electron microscope (SEM) techniques.<br />

6.2 X-ray fluorescence (XRF) studies<br />

6.2.1 Scope <strong>an</strong>d method<br />

Chemical <strong>an</strong>alyses were carried out using X-ray fluorescence technique to determine the<br />

percentage composition <strong>of</strong> major oxides in the clay soils, <strong>an</strong>d include SiO2, Al2O3, Fe2O3,<br />

FeO, MgO, CaO, MnO, Na2O, K2O, TiO2, P2O5 <strong>an</strong>d SO4. Percentage content <strong>of</strong> F, Cl, S<br />

<strong>an</strong>d Cr, as well as losses on ignition were also determined. The <strong>an</strong>alysis results have been<br />

useful in computing mineral contents in the tested soils (according to Olphen <strong>an</strong>d Fripiat,<br />

1979), especially where direct measurement <strong>of</strong> mineralogical compositions, such as kaolinite<br />

in red soils, has proved difficult.<br />

A PW-1480 (Rh 100kV LiF220 Ge111 T1AP) X-ray spectrometer <strong>an</strong>d a shale (<strong>clays</strong>tone)<br />

st<strong>an</strong>dard were employed in the XRF <strong>an</strong>alysis. Preparation <strong>of</strong> test specimens involved mixing<br />

<strong>of</strong> Hoechst wax (as diluent) with soil sample at a ratio <strong>of</strong> 1:5 to make a homogenous mixture.<br />

6.2.2 Results<br />

Results <strong>of</strong> chemical <strong>an</strong>alyses obtained for the soils in this study are presented in Table 6.1 for<br />

red soils <strong>an</strong>d Table 6.2 for black <strong>clays</strong>.<br />

Table 6.1. Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> red soils obtained in this study.<br />

% content <strong>of</strong>: Rd1-30cm Rd1-100cm Rd1-200cm Rd1-400cm<br />

SiO2 47,50 47,10 47,20 47,6<br />

Al2O3 32,20 33,40 33,20 33,30<br />

Fe2O3 15,30 15,50 15,50 15,20<br />

FeO - - - -<br />

BaO 0,062 0,061 0,071 0,064<br />

MgO 0,12 0,097 0,14 0,13<br />

CaO 0,26 0,24 0,16 0,26<br />

Na2O ‹ ‹ ‹ ‹<br />

K2O 0,96 0,66 0,68 0,60<br />

P2O5 0,19 0,086 0,096 0,062<br />

ZrO2 0,37 0,37 0,37 0,36<br />

TiO2 1,57 1,51 1,51 1,46<br />

MnO 0,96 0,53 0,61 0,52<br />

Loss on ignition - - - -<br />

Total 99,49 99,55 99,54 99,56<br />

Rd1-30cm, Rd1-100cm, Rd1-200cm & Rd1-400cm are red soil samples.


59<br />

The presence <strong>of</strong> BaO in the red <strong>an</strong>d black soils is most likely a result <strong>of</strong> contributions in the<br />

form <strong>of</strong> BaCO3 <strong>an</strong>d/ or BaSO4 as contamin<strong>an</strong>ts from industrial effluents <strong>an</strong>d/ or wastes.<br />

Fertilizers <strong>an</strong>d industrial water could also account for the presence <strong>of</strong> traces <strong>of</strong> P2O5 in both<br />

types <strong>of</strong> soil.<br />

Table 6.2. Results <strong>of</strong> chemical <strong>an</strong>alyses <strong>of</strong> black <strong>clays</strong> obtained in this study.<br />

Content SA2- SA2- SA41- SB1- SB1- SB41- SB41- SC17- SC41- SC41-<br />

<strong>of</strong> (%): 70cm 105cm 50cm 50cm 70cm 30cm 50cm 50cm 30cm 50cm<br />

SiO2 52,50 51,40 58,35 52,79 53,44 52,64 53,21 56,38 52,06 53,71<br />

Al2O3 14,71 14,93 13,37 12,74 12,84 12,14 12,20 11,87 12,06 11,77<br />

Fe2O3 8,66 8,85 6,98 6,90 6,96 6,67 6,75 6,20 6,49 6,10<br />

FeO - - - - - - - - - -<br />

BaO 0,052 0,052 0,031 0,062 0,062 0,064 0,062 0,085 0,123 0,104<br />

MgO 1,25 1,22 1,13 1,16 1,23 1,43 1,42 1,28 1,34 1,31<br />

CaO 1,28 4,29 1,28 1,54 2,01 2,61 2,32 1,47 1,64 1,48<br />

Na2O 1,06 1,04 0,69 0,71 0,76 0,50 0,56 0,82 0,90 0,89<br />

K2O 1,78 1,73 1,25 1,27 1,34 1,02 1,04 1,41 1,32 1,31<br />

P2O5 0,026 0,024 0,037 0,032 0,038 0,038 0,038 0,047 0,036 0,04<br />

ZrO2 0,092 0,091 0,085 0,091 0,092 0,078 0,078 0,072 0,073 0,072<br />

TiO2 0,96 0,95 0,90 0,76 0,75 0,77 0,77 0,75 0,79 0,77<br />

MnO 0,312 0,363 0,249 0,455 0,439 0,497 0,505 0,488 0,469 0,511<br />

SO3 - - - - - - - 0,011 0,028 0,016<br />

Cl 0,065 0,039 - - - - - - 0,015 0,015<br />

F 0,262 0,362 0,247 0,294 0,325 0,281 0,267 0,311 0,300 0,293<br />

Loss on - - - - - - - - - -<br />

ignition<br />

Total 83,01 85,34 84,60 78,80 80,29 78,74 79,22 81,19 77,64 78,39<br />

SA2-70cm, SA2-105cm, SA41-50cm, SB1-50cm, SB1-70cm, SB41-30cm, SB41-50cm, SC17-50cm, SC41-<br />

30cm <strong>an</strong>d SC41-50cm are black clay samples.<br />

The total sum <strong>of</strong> the various chemical components <strong>of</strong> the black <strong>clays</strong> add up to only 77,64 –<br />

85,34% (Table 6.2), <strong>an</strong>d not 100%. This is a relatively large difference which could be partly<br />

accounted for by the presence in the soils <strong>of</strong> accessory amounts <strong>of</strong> heavy metal compounds,<br />

light <strong>an</strong>d noble elements; <strong>an</strong>d partly by the loss on ignition <strong>of</strong> org<strong>an</strong>ic matter components. The<br />

difference is comparatively small <strong>an</strong>d negligible in the red soils (Table 6.1) in which the total<br />

sum <strong>of</strong> the chemical components is virtually 100%, probably due to the relatively low <strong>an</strong>d/ or<br />

negligible org<strong>an</strong>ic content (Table 6.7).<br />

6.3 X-ray diffraction (XRD) studies<br />

6.3.1 Scope <strong>an</strong>d method<br />

X-ray diffraction studies <strong>an</strong>d <strong>an</strong>alyses were carried out in the present work to determine the<br />

mineralogical composition <strong>of</strong> the black <strong>clays</strong> <strong>an</strong>d red soils. Clay minerals present were<br />

identified by their characteristic diffraction patterns.<br />

A diffractometer (Type PW1710-Basis) was used to determine the presence <strong>of</strong> clay minerals<br />

which include smectites ( montmorillonite), kaolinite <strong>an</strong>d illite. This was done for selected<br />

conditions <strong>of</strong> continuous sc<strong>an</strong>ning rate <strong>of</strong> 0,020 (° 2θ/second), chart speed 0,10 (mm/ second)<br />

<strong>an</strong>d using Cu-Kα radiation. The clay minerals were X-rayed from prepared clay fractions (


60<br />

2µm size) filtered both by me<strong>an</strong>s <strong>of</strong> a programmable centrifuge as well as through a filter<br />

membr<strong>an</strong>e. The original soil sample had been previously <strong>an</strong>d moderately ground to reduce<br />

strong textural effects usually caused by presence <strong>of</strong> quartz <strong>an</strong>d feldspars. Treatment with<br />

hydrogen peroxide (3-10% H2O2) to oxidise <strong>an</strong>y org<strong>an</strong>ic matter found had also been done,<br />

followed by treatment with ammonia solution (0,01 N NH3-water) to remove the carbonate<br />

components <strong>an</strong>d prevent Ca ions from flocculating the clay particles in prepared suspensions.<br />

Calcium fluoride st<strong>an</strong>dard was adopted where upon preparations <strong>of</strong> 200mg CaF2 mixed with<br />

1g <strong>of</strong> soil sample were used.<br />

The presence <strong>of</strong> swelling clay minerals (smectites) was investigated by measuring <strong>an</strong>d<br />

comparing diffractograms from filtered air-dried specimens with those from filtered <strong>an</strong>d<br />

glycolised specimens. Glycolisation involved heating specimens to 40 °C to collapse the<br />

layers <strong>of</strong> <strong>an</strong>y montmorillonite present, followed by heating in a chamber <strong>of</strong> glycol<br />

atmosphere. Usually glycol has the same effect as water <strong>an</strong>d, in the presence <strong>of</strong> smectites,<br />

would go in-between the layers <strong>an</strong>d move them apart, thereby causing pronounced exp<strong>an</strong>sion.<br />

The difference in the amount <strong>of</strong> impulse (reflex intensity) between the resulting<br />

diffractograms <strong>an</strong>d / or smectite peaks <strong>an</strong>d those from air-dried specimens would therefore be<br />

signific<strong>an</strong>t (Reynolds <strong>an</strong>d Moore, 1989; Sattler, 2000).<br />

The clay minerals were identified from basal reflections obtained from their sheet-like<br />

structure <strong>an</strong>d by matching the obtained diffraction patterns with calculated one – dimensional<br />

diffraction pr<strong>of</strong>iles ( Reynolds <strong>an</strong>d Hower, 1970; Reynolds <strong>an</strong>d Moore, 1989; Sattler, 2000).<br />

Primary non-clay minerals were also identified by sc<strong>an</strong>ning the preparations at selected<br />

r<strong>an</strong>ges; <strong>an</strong>d include quartz, calcite, dolomite <strong>an</strong>d iron oxides. The minerals were identified by<br />

matching their characteristic diffraction peaks with known st<strong>an</strong>dards.<br />

6.3.2 Results<br />

The results <strong>of</strong> percentage composition <strong>of</strong> clay minerals in the black <strong>clays</strong> are summarised in<br />

Table 6.3, while the corresponding diffractograms are given in Figures 6.1 to 6.6. The<br />

diffractograms obtained for the red soils are presented in Figures 6.7 to 6.10. However, it was<br />

not possible to determine the percentage clay mineral composition <strong>of</strong> the red soils. This is<br />

most probably due to the high contents <strong>of</strong> iron in these soils which caused formation <strong>of</strong> very<br />

thin layers <strong>of</strong> specimens on glass slides, thus making it difficult to prepare oriented samples<br />

necessary for drop <strong>an</strong>alysis <strong>an</strong>d/ or pipette <strong>an</strong>alysis methods.<br />

The results <strong>of</strong> the <strong>an</strong>alyses show the black <strong>clays</strong> to be composed <strong>of</strong> over 90% smectites<br />

(montmorillonite: CaNaMgFeAlSiOOH.HO) <strong>an</strong>d less th<strong>an</strong> 10% kaolinite (Al2Si2O5(OH)4).<br />

Illite (K-Na-Mg-Fe-Al-Si-O-H2O) occasionally occurs in trace form, <strong>an</strong>d is probably <strong>an</strong><br />

alteration product <strong>of</strong> the feldspars <strong>an</strong>d / or kaolinite. Also found contained are quartz (4-9%)<br />

<strong>an</strong>d accessories <strong>of</strong> K-feldspars [s<strong>an</strong>idine (K,Na)AlSi3O8), orthoclase (Na, K)Si3O8),<br />

microcline (KAlSi3O8)], haematite (Fe2O3) <strong>an</strong>d carbonates, i.e.calcite (CaCO3), dolomite<br />

(CaMg(CO3)2, <strong>an</strong>kerite (Ca(Fe,Mg)(CO3)2, siderite (FeCO3)].<br />

According to Coduto (1994), montmorillonite <strong>of</strong>ten results from weathering <strong>of</strong><br />

ferromagnesi<strong>an</strong> minerals, calcic feldspars <strong>an</strong>d volc<strong>an</strong>ic materials. In addition, sodium<br />

montmorillonite is <strong>of</strong>ten formed from weathering <strong>of</strong> volc<strong>an</strong>ic ash, while other<br />

montmorillonites form in environments <strong>of</strong> alkaline conditions characterised by a supply <strong>of</strong><br />

magnesium ions <strong>an</strong>d a lack <strong>of</strong> leaching. The montmorillonite in black <strong>clays</strong> <strong>of</strong> the present<br />

study must have resulted from weathering <strong>an</strong>d alteration <strong>of</strong> volc<strong>an</strong>ic ash previously deposited<br />

in a basin-like lake environment. Additional contribution in montmorillonite formation


61<br />

resulted most probably from supplied magnesium <strong>an</strong>d other cations leached down from<br />

surrounding higher areas <strong>an</strong>d deposited under conditions <strong>of</strong> impeded drainage <strong>an</strong>d high<br />

alkalinity in the basin. The alkalinity (contributed by seepage waters containing soluble bases<br />

<strong>of</strong> Mg, Ca, Na, K leached down from higher areas <strong>of</strong> red soils) served to facilitate the Si/Al<br />

build-up <strong>an</strong>d properties characteristic <strong>of</strong> montmorillonite. The impeded drainage further<br />

serves to retain most <strong>of</strong> the above soluble bases necessary in the build-up <strong>of</strong> montmorillonite<br />

structure. Presence <strong>of</strong> accessories <strong>of</strong> haematite could be attributed to tr<strong>an</strong>sported components<br />

from areas <strong>of</strong> red soils.<br />

Table 6.3. Clay mineral composition <strong>of</strong> black <strong>clays</strong>.<br />

Sample No. Smectites<br />

(montmorillonite)<br />

(%)<br />

Il lite<br />

(%)<br />

Kaolinite<br />

(%)<br />

Quartz<br />

(%)<br />

K-feldspars<br />

(s<strong>an</strong>idine,<br />

orthoclase)<br />

SA2-70cm >90 - 90 trace 95 - 95 - 95 - 95 trace 95 - 95 - 95 - 95 trace


62<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

SA 2 - 105cm + St<strong>an</strong>dard<br />

Q: Quartz<br />

Sm: Smectites<br />

K: Kaolinite<br />

Or: Orthoclase<br />

Sn: S<strong>an</strong>idine<br />

Mc: Microcline<br />

H: Haematite<br />

Cc: Calcite<br />

Ak: Ankerite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

Sn<br />

F<br />

232<br />

F<br />

F<br />

400<br />

300<br />

200<br />

100<br />

Quartz: 4%<br />

Calcite: 3% (s.a. Scheibler)<br />

Accessories: K - Feldspar, Haematite,<br />

Dolomite, Ankerite<br />

Rest: Clay minerals<br />

(S. Texture preparation)<br />

Sm<br />

Mc<br />

K Sn Sn<br />

57<br />

K<br />

K Or<br />

Sm<br />

Q Sn<br />

Sn<br />

Sn<br />

Or<br />

Mc<br />

Q<br />

Cc<br />

32<br />

Sn<br />

Tr<br />

K<br />

Sn<br />

Cc<br />

Sm Cc<br />

Q K H Sn<br />

Q<br />

Cc QAk<br />

Q<br />

H<br />

Cc<br />

Q<br />

Sm<br />

H<br />

Cc<br />

Q<br />

F<br />

0.0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.2. X-ray diffraction diagram <strong>of</strong> black <strong>clays</strong> collected at 1,05m depth.<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

SB 1 -70cm + St<strong>an</strong>dard<br />

Q: Quartz<br />

Sm: Smectites<br />

I: Illite<br />

K: Kaolinite<br />

Or: Orthoclase<br />

Sn: S<strong>an</strong>idine<br />

Mc: Microcline<br />

H: Haematite<br />

Sd: Siderite<br />

Ak: Ankerite<br />

D: Dolomite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

Mc<br />

Sn<br />

F<br />

275<br />

F<br />

F<br />

400<br />

300<br />

Quartz: 4%<br />

Accessories: K -Feldspars, Haematite<br />

Carbonates<br />

Rest: Clay minerals<br />

(S. Texture preparation)<br />

I<br />

Q<br />

Tr<br />

200<br />

100<br />

Sm<br />

I<br />

K<br />

Mc<br />

Sn Sn<br />

I<br />

K<br />

Sm<br />

56<br />

Or Or<br />

Sn K<br />

D<br />

Sn<br />

Sn<br />

Sm<br />

Q H Or<br />

Sd Q Sn<br />

Q<br />

Sd<br />

Ak<br />

Q<br />

Or<br />

Sm<br />

Q<br />

D<br />

H<br />

Q<br />

0.0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.3. X-ray diffraction diagram <strong>of</strong> black <strong>clays</strong> collected at 0,70m depth.


63<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

SC 17 -50cm + St<strong>an</strong>dard<br />

Q: Quartz Td: Tridymite<br />

Sm: Smectites<br />

K: Kaolinite Cr: Cristobalite<br />

Or: Orthoclase<br />

Sn: S<strong>an</strong>idine<br />

Mc: Microcline<br />

H: Haematite<br />

Sd: Siderite<br />

Ak: Ankerite<br />

D: Dolomite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

F<br />

256<br />

F<br />

F<br />

400<br />

300<br />

Quartz: 4%<br />

Accessories: K -Feldspars, Haematite,<br />

Carbonates<br />

Rest: Clay minerals<br />

Q<br />

(S. Texture preparation)<br />

Sn<br />

66 Mc<br />

Tr<br />

200<br />

100<br />

Sm<br />

Or<br />

K SnOr<br />

K<br />

Q<br />

Sm Or<br />

Q Sn<br />

Sm<br />

Td Sn<br />

Cr Or Mc<br />

K<br />

Sn<br />

Sd Sn<br />

H Q Q<br />

Q<br />

Sd<br />

Ak<br />

Q<br />

H<br />

Sm<br />

Q<br />

D<br />

H<br />

Q<br />

0.0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.4 X-ray diffraction diagram <strong>of</strong> yellow-brown to black <strong>clays</strong> collected at 0,50m depth.<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

SC 33 -30cm + St<strong>an</strong>dard<br />

Q: Quartz Td: Tridymite<br />

Sm: Smectites Cr: Cristobalite<br />

Or: Orthoclase<br />

Sn: S<strong>an</strong>idine<br />

Mc: Microcline<br />

H: Haematite<br />

Sd: Siderite<br />

Ak: Ankerite<br />

D: Dolomite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

Q<br />

F<br />

252<br />

F<br />

F<br />

400<br />

Quartz: 9%<br />

Accessories: K -Feldspars, Haematite,<br />

Mc<br />

Carbonates<br />

I Sn<br />

Rest: Clay minerals<br />

(S. Texture preparation)<br />

Tr<br />

300<br />

125<br />

200<br />

100<br />

Sm<br />

Or<br />

Or<br />

Sn Sn<br />

Sm Or<br />

Q<br />

Sn TdCr<br />

Sn<br />

Or<br />

Ak<br />

Sm<br />

Q H Sd<br />

Q Q<br />

Sd Sn<br />

Sn<br />

D<br />

Q<br />

Sd<br />

Ak<br />

Q<br />

AkSm<br />

D<br />

Q D<br />

H H<br />

Q<br />

Sd<br />

0.0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.5. X-ray diffraction diagram <strong>of</strong> black <strong>clays</strong> collected at 0,30m depth.


64<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

SC 33 -50cm + St<strong>an</strong>dard<br />

Q: Quartz Td: Tridymite<br />

Sm: Smectites Cr: Cristobalite<br />

Or: Orthoclase<br />

Sn: S<strong>an</strong>idine<br />

Mc: Microcline<br />

H: Haematite<br />

Sd: Siderite<br />

Ak: Ankerite<br />

D: Dolomite<br />

Q<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

F<br />

254<br />

F<br />

F<br />

400<br />

Quartz: 7%<br />

Accessories: K -Feldspars, Haematite,<br />

Carbonates<br />

Rest: Clay minerals<br />

(S. Texture preparation)<br />

300<br />

200<br />

100<br />

Sm<br />

Or Or<br />

Sn Sn<br />

Sm<br />

Q<br />

Td<br />

Cr<br />

Or<br />

98 Sn<br />

Sn Mc<br />

Mc<br />

Sm<br />

Ak Q Sd<br />

D<br />

Sd Q<br />

Or<br />

Sn<br />

H Q Sn<br />

Mc<br />

Tr<br />

Mc<br />

Q<br />

Q<br />

Sd<br />

Ak<br />

Q<br />

H<br />

Sm<br />

Q<br />

D<br />

H<br />

Q<br />

0.0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.6. X-ray diffraction diagram <strong>of</strong> black <strong>clays</strong> collected at 0,50m depth.<br />

The diffractograms for the red soils show them as containing mainly kaolinite <strong>an</strong>d haematite<br />

with accessories <strong>of</strong> quartz. According to Day (2001) <strong>an</strong>d Holtz & Kovacs (1981), kaolinite<br />

belongs to kaolin minerals, i.e. a group <strong>of</strong> clay minerals consisting <strong>of</strong> hydrous aluminium<br />

silicates. Approximate mineralogical compositions <strong>of</strong> red soils could therefore be estimated<br />

from results <strong>of</strong> chemical <strong>an</strong>alyses (Table 6.1) obtained for the soils. This was done by<br />

assuming that the kaolinite portion is constituted mainly by SiO2 <strong>an</strong>d Al2O3; <strong>an</strong>d haematite<br />

by Fe2O3. The resulting estimated mineralogical compositions for the red soils are presented<br />

in Table 6.4, <strong>an</strong>d show the red soils as being mainly kaolinite (80-81%) <strong>an</strong>d haematite (15-<br />

16%) sediments; with accessories <strong>of</strong> quartz (1-3%) as well as minerals <strong>of</strong> tit<strong>an</strong>ium (about 2%)<br />

<strong>an</strong>d m<strong>an</strong>g<strong>an</strong>ese (about 1%). The kaolinite must have resulted from weathering <strong>an</strong>d alteration<br />

<strong>of</strong> aluminiumsilicate-rich feldspars in the underlying Nairobi trachytes <strong>an</strong>d other volc<strong>an</strong>ic<br />

materials, under humid conditions (rainfall) <strong>an</strong>d in the presence <strong>of</strong> carbon dioxide gas.<br />

Table 6.4. Estimated mineralogical composition <strong>of</strong> red soils.<br />

Sample No. Kaolinite<br />

(%)<br />

Haematite<br />

(%)<br />

Quartz<br />

(%)<br />

Tit<strong>an</strong>ium<br />

(TiO2)<br />

M<strong>an</strong>g<strong>an</strong>ese<br />

(MnO2)<br />

(%)<br />

Rd1-30cm 80 15 1 2 1<br />

Rd1-100cm 81 16 3 2 1<br />

Rd1-200cm 80 16 2 2 1<br />

Rd1-400cm 81 15 1 2 1<br />

In addition, hydrothermal alteration <strong>of</strong> the aluminiumsilicates could have formed<br />

pseudomorphs <strong>of</strong> kaolinite from feldspars <strong>an</strong>d/or muscovite, topas, leucite, <strong>an</strong>dalusite, <strong>an</strong>d<br />

pyrophyllite contained in the rocks. Kaolinite-rich red soils containing alkali <strong>an</strong>d alkali-earth<br />

metals (Na, K, Mg, Ca) usually alter during favourable conditions into secondary feldspars,


65<br />

sericite <strong>an</strong>d chlorite (see next section on sc<strong>an</strong>ning electron microscope studies); while those<br />

deficient in the alkali <strong>an</strong>d alkali-earth metals alter into aluminiumsilicates <strong>of</strong> <strong>an</strong>dalusite,<br />

sillim<strong>an</strong>ite <strong>an</strong>d ky<strong>an</strong>ite.<br />

Theoretical chemical compositions <strong>of</strong> the minerals kaolinite, illite <strong>an</strong>d montmorillonite are<br />

given in Table 6.5. A comparison <strong>of</strong> the results in this table with those <strong>of</strong> chemical <strong>an</strong>alyses in<br />

Table 6.1 suggest the kaolinite in red soils to be depleted in Al2O3. The depletion is most<br />

probably a result <strong>of</strong> low PH conditions caused by formation <strong>of</strong> weak carbonic acids in the<br />

soils through org<strong>an</strong>ic matter decomposition <strong>an</strong>d/ or hydrothermal alteration at signific<strong>an</strong>t<br />

temperatures.<br />

Table 6.5. Theoretical compositions <strong>of</strong> kaolinite, illite <strong>an</strong>d montmorillonite<br />

(Jasmund & Lagaly, 1993; Moore & Reynolds, 1989; Rösler, 1980).<br />

% content <strong>of</strong>: Smectites<br />

Il lite Kaolinite<br />

(montmorillonite)<br />

SiO2 48-56 51-54 46,55<br />

Al2O3 11-22 21-29 39,5<br />

H2O 12-24 - 10,0-13,95<br />

Fe2O3 ≥5 0,5-5,3 Trace<br />

FeO - 0,9-1,2 -<br />

BaO - - Trace<br />

MgO 4-9 2,8-3,5 Trace<br />

CaO 0,8-3,3 ,02-,05 Trace<br />

Na2O trace ,08-,10 Trace<br />

K2O trace 2-9 Trace<br />

P2O5 - 0,6-0,82 -<br />

ZrO2 - - -<br />

TiO2 - 0,60-0,82 Trace<br />

MnO - - -<br />

Loss on ignition - - -<br />

Total >80 >79,5 96,0-100,0<br />

800<br />

Impulse<br />

RD 1 - 30cm + St<strong>an</strong>dard<br />

F<br />

F<br />

700<br />

600<br />

Q: Quartz (1%)<br />

H: Haematite<br />

K: Kaolinite<br />

F: Fluorite (CaF2- St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

500<br />

F<br />

400<br />

214<br />

300<br />

Tr<br />

200<br />

100<br />

K<br />

K<br />

Q<br />

Q<br />

18<br />

K<br />

H<br />

H<br />

K<br />

Q<br />

K<br />

H<br />

Q<br />

H Q<br />

H<br />

H<br />

Q<br />

H<br />

K<br />

H<br />

Q<br />

F<br />

0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.7. X-ray diffraction diagram <strong>of</strong> red soils collected at 0,30m depth.


66<br />

800<br />

Impulse<br />

700<br />

600<br />

RD 1 - 100cm + St<strong>an</strong>dard<br />

Q: Quartz (3%)<br />

H: Haematite<br />

K: Kaolinite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

F<br />

216<br />

F<br />

500<br />

F<br />

400<br />

Tr<br />

300<br />

200<br />

100<br />

K<br />

K K Q<br />

Q<br />

35<br />

K<br />

H<br />

H<br />

H<br />

K<br />

K<br />

Q<br />

K<br />

H<br />

Q<br />

H<br />

H<br />

K<br />

Q<br />

H<br />

K<br />

H<br />

F<br />

Q<br />

0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.8. X-ray diffraction diagram <strong>of</strong> red soils collected at 1,0m depth.<br />

800<br />

Impulse<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

RD 1 - 200cm + St<strong>an</strong>dard<br />

Q: Quartz (2%)<br />

H: Haematite<br />

K: Kaolinite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

K<br />

Q H<br />

Q<br />

24<br />

F<br />

211<br />

H<br />

K<br />

H K<br />

Q K<br />

H<br />

Tr<br />

F<br />

H Q<br />

H<br />

F<br />

Q<br />

H<br />

K<br />

H<br />

F<br />

H<br />

0<br />

0 20 40 60 [ ° 2 ]<br />

Figure 6.9. X-ray diffraction diagram <strong>of</strong> red soils collected at 2,0m depth.


67<br />

Q: Quartz (1%)<br />

H: Haematite<br />

K: Kaolinite<br />

F: Fluorite (CaF2-St<strong>an</strong>dard)<br />

Tr: Specimen holder<br />

204<br />

16<br />

H<br />

0.0<br />

Figure 6.10. X-ray diffraction diagram <strong>of</strong> red soils collected at 4,0m depth.<br />

6.4 Sc<strong>an</strong>ning electron microscope (SEM) <strong>an</strong>alysis<br />

6.4.1 Scope <strong>an</strong>d method<br />

The sc<strong>an</strong>ning electron microscope technique was applied on fresh undisturbed soil specimens<br />

sliced into suitable volume (2cm*2cm*2cm). The <strong>an</strong>alysis was also performed on prepared<br />

specimens <strong>of</strong> the underlying phonolitic <strong>an</strong>d trachytic rocks. The <strong>an</strong>alysis involved establishing<br />

the morphology, mode <strong>of</strong> occurrence relative to clayey matrix <strong>an</strong>d diagenetic features <strong>of</strong> clay<br />

<strong>an</strong>d other minerals. Possible alterations <strong>of</strong> clay minerals to form other minerals (<strong>an</strong>d vice<br />

versa) were also sought <strong>an</strong>d identified; while relative abund<strong>an</strong>ce <strong>an</strong>d type <strong>of</strong> texture <strong>an</strong>d/ or<br />

fabric (flocculation, r<strong>an</strong>domness, orientations, lamination) <strong>of</strong> minerals were also investigated<br />

<strong>an</strong>d established. In addition, the porosity, interconnection <strong>of</strong> pores <strong>an</strong>d permeability<br />

(possibility <strong>of</strong> water movement/ percolation) <strong>of</strong> the soils were assessed.<br />

The <strong>an</strong>alysis employed a Leiz Sc<strong>an</strong>ning electron microscope (Modell: ISI Super 40/3A)<br />

m<strong>an</strong>ufactured by International Scientific Instruments (ISI, 1972). The microscope<br />

incorporated a Eumex detector having Lithium fitted with <strong>an</strong> ultra-thin Beryllium window to<br />

facilitate element measurements in the r<strong>an</strong>ge <strong>of</strong> carbon (C) to Ur<strong>an</strong>ium (U). Use was also<br />

made <strong>of</strong> Hardware for High Perform<strong>an</strong>ce X-ray Micro-<strong>an</strong>alysis (EDX: WINEDS) as well as<br />

Wineds 3.0 & Edisson 32 S<strong>of</strong>tware for element <strong>an</strong>alysis (Thomson Scientific Instruments Pty<br />

Ltd. & Fa.Getac, 1995). In addition, Hardware for digital photo-processing with<br />

Rasterelektronenmikroskop (REM PC Einsteckkarte); as well as WinDISS 2.0 S<strong>of</strong>tware for<br />

digital photo-taking, were employed (Fa.Point Elektronic GmbH, 1997).


68<br />

6.4.2 Results<br />

6.4.2.1 Nairobi phonolite<br />

Results <strong>of</strong> sc<strong>an</strong>ning electron microscope <strong>an</strong>alysis show Nairobi phonolite to exhibit a<br />

generally massive, homogeneous <strong>an</strong>d compact structure (Plate 6.1). Texture is mainly <strong>of</strong> large<br />

phenocrysts embedded in a fine-grained matrix (Plate 6.3). The phenocrysts show presence <strong>of</strong><br />

O, Al, Si, K implying K-feldspars; <strong>an</strong>d some Fe <strong>an</strong>d Ti probably due to inclusions <strong>of</strong> iron <strong>an</strong>d<br />

tit<strong>an</strong>ium ore. The phenocrysts show partly corroded <strong>an</strong>d/ or leached <strong>an</strong>d altered surfaces<br />

(Plate 6.2). The matrix also shows <strong>an</strong> increased presence <strong>of</strong> Si, K <strong>an</strong>d Fe, as well as some O<br />

<strong>an</strong>d Al, implying a predomin<strong>an</strong>ce <strong>of</strong> K-feldspars with some iron oxides <strong>an</strong>d/ or ores. The iron<br />

oxides are most likely a result <strong>of</strong> weathering <strong>an</strong>d/ or chemical decomposition <strong>of</strong> phonolite to<br />

form tuffs as well as ferricrete <strong>an</strong>d / or iron concretions (Plates 6.4/6.5). Occasional presence<br />

<strong>of</strong> Ca in the matrix could be attributed to weathering <strong>of</strong> phonolitic tuff to form secondary<br />

limestone <strong>an</strong>d/or calcite (Plate 6.5).<br />

Plate 6.1. Massive <strong>an</strong>d compact structure<br />

Plate 6.2. Corroded, leached K-feldspar


69<br />

Plate 6.3. K-feldspar phenocryst in fine matrix (diagram: K-feldspar phenocryst).<br />

Plate 6.4. Alteration <strong>of</strong> rock forming ferricrete<br />

(diagram: composition <strong>of</strong> matrix).<br />

Plate 6.5.Alteration <strong>of</strong> rock/tuff forming<br />

ferricrete <strong>an</strong>d secondary calcite.<br />

(diagram: matrix).


70<br />

6.4.2.2 Nairobi trachytes<br />

Plate 6.6. Less compact structure formed by<br />

plates <strong>of</strong> fine K-feldspar phenocrysts.<br />

Plate 6.7. Relatively more porous structure<br />

formed by tabular K-feldspar crystals.<br />

Plate 6.8. Large tabular K-feldspar phenocrysts,<br />

with Fe <strong>an</strong>d Ti-rich crusts as part <strong>of</strong> matrix.<br />

Plate 6.9. K-feldspar crystals with leaching<br />

surfaces <strong>an</strong>d solution pores.


71<br />

The trachytes generally show a relatively less compact <strong>an</strong>d more porous structure (Plates 6.6/<br />

6.7), <strong>an</strong>d are characterised by plates <strong>of</strong> tabular K-feldspar phenocrysts showing patched<br />

overgrowths <strong>of</strong> Fe <strong>an</strong>d Ca-rich crusts (Plate 6.8). The phenocrysts show a composition <strong>of</strong> O,<br />

Al, Si, K <strong>an</strong>d occasionally Na, <strong>an</strong>d this is suggestive <strong>of</strong> s<strong>an</strong>idine. The feldspars show effects<br />

<strong>of</strong> weathering <strong>an</strong>d alteration through leached surfaces <strong>an</strong>d solution pores (Plate 6.9). The<br />

matrix is predominated by K-feldspars showing presence <strong>of</strong> O, Al, Si, K suggestive <strong>of</strong><br />

s<strong>an</strong>idine <strong>an</strong>d/or orthoclase; while crusts rich in Fe <strong>an</strong>d Ca also occur as weathering <strong>an</strong>d<br />

alteration products <strong>of</strong> the rock.<br />

6.4.2.3 Black <strong>clays</strong><br />

The black <strong>clays</strong> exhibit a generally heavy, dense <strong>an</strong>d massive appear<strong>an</strong>ce with solution pores,<br />

cavities <strong>an</strong>d / or cracks found within them (Plates 6.10, 6.11 & 6.24). Quartz grains <strong>an</strong>d K-<br />

feldspar crystals show overgrowths <strong>of</strong> Fe, Ca <strong>an</strong>d Mg-rich crusts (Plates 6.12, 6.15); <strong>an</strong>d<br />

occur in a fine, weathered <strong>an</strong>d generally smectitic clay matrix (Plate 6.23). The generally<br />

smooth sub-rounded quartz grains (Plates 6.19, 6.20) indicate water tr<strong>an</strong>sport, <strong>an</strong>d this serves<br />

to support the hypothesis that part <strong>of</strong> the materials forming black <strong>clays</strong> have been contributed<br />

by erosional processes from the surrounding high areas. The feldspars show alteration effects<br />

through corroded <strong>an</strong>d leached solution surfaces (Plate 6.18). Apparent detrital components<br />

showing presence <strong>of</strong> O, Ca, Fe, K, Si, Ti also occur <strong>an</strong>d suggest biotite (Plate 6.14), a<br />

probable result <strong>of</strong> weathered phonolitic <strong>an</strong>d/ or metamorphic rock inclusions. Iron concretions<br />

(Plates 6.17, 6.21) <strong>an</strong>d blocks <strong>of</strong> <strong>an</strong>kerite (Plate 6.22) showing elevated Fe content are also<br />

found embedded in the clay matrix. Traces <strong>of</strong> org<strong>an</strong>ic material (Plate 6.16) are also found.<br />

Occasional weathered plates <strong>of</strong> kaolinite also form part <strong>of</strong> the clay matrix (Plate 6.13). The<br />

presence <strong>of</strong> Fe <strong>an</strong>d Ca-rich crusts could be attributed to underlying phonolitic rocks<br />

weathering through tuff <strong>an</strong>d ferricrete thereby forming secondary limestone <strong>an</strong>d iron<br />

concretions (murram); <strong>an</strong>d partly to weathering <strong>of</strong> the smectites in the clay matrix. The<br />

solution pores are a probable result <strong>of</strong> the iron, limestone <strong>an</strong>d/ or other carbonate components<br />

going into reaction/ solution, while the partings <strong>an</strong>d cracks are most likely due to shrinkage<br />

effects in the smectitic clay matrix in a dry condition. Presence <strong>of</strong> Ti <strong>an</strong>d Zn in the matrix<br />

could be attributed to heavy mineral inclusions tr<strong>an</strong>sported from surrounding areas.<br />

Plate 6.10. Massive, dense, heavy <strong>clays</strong> with<br />

solution pores.<br />

Plate 6.11. Partings, cracks due to shrinkage;<br />

<strong>an</strong>d solution pores.


72<br />

Plate 6.12 Quartz grain with overgrowth <strong>of</strong><br />

Fe, Mg <strong>an</strong>d Ca-rich crusts.<br />

Plate 6.13. Weathered kaolinite (K) in<br />

smectitic clay matrix with Fe <strong>an</strong>d Ca crusts.<br />

Plate 6.14. Inclusions <strong>of</strong> detrital biotite in<br />

weathered clay matrix.<br />

Plate 6.15. K-feldspar phenocryst in smectitic<br />

clay matrix.


73<br />

Plate 6.16. Org<strong>an</strong>ic matter (remains <strong>of</strong> pl<strong>an</strong>t roots) in a weathering smectitic clay matrix.<br />

Plate 6.17. Iron concretion. Al, Si, K, Ti due<br />

to contribution from clay matrix.<br />

Plate 6.18. Corroded/ leached K-feldspar.<br />

Fe, Ti <strong>an</strong>d Zn due to clay matrix.


74<br />

Plate 6.19. Sub-rounded quartz grains in<br />

weathering smectitic matrix.<br />

Plate 6.20. Sub-rounded quartz grain in clay<br />

matrix.<br />

Plate 6.21. Iron concretions (o) in smectitic<br />

clay matrix (s).<br />

Plate 6.22. Blocky <strong>an</strong>kerite in clay matrix.<br />

Plate 6.23. Smectitic clay matrix exhibiting Al, Ca, Fe, <strong>an</strong>d high Si content. Crusts with K,<br />

Ti, Zn found associated.


75<br />

Plate 6.24. Solution pores/ cavities in weathering smectitic clay matrix with Al, Ca, Fe,<br />

<strong>an</strong>d high Si content; <strong>an</strong>d associated K, Ti crusts.<br />

6.4.2.4 Red soils<br />

The red soils show a relatively loose, porous <strong>an</strong>d friable structure <strong>of</strong> weakly bound or<br />

cemented lumps <strong>of</strong> clay matrix (Plate 6.25). Large grains <strong>of</strong> iron concretions <strong>an</strong>d quartz are<br />

embedded in a fine weathering clay matrix. The iron concretions show <strong>an</strong> elevated Fe content,<br />

up to 70% by weight (Plate 6.29); <strong>an</strong>d weather through leaching to form solution pores (Plate<br />

6.30). Quartz grains are slightly abraded <strong>an</strong>d rounded to sub-rounded (Plate 6.28), <strong>an</strong>d are<br />

probably detrital, having been derived by erosion <strong>an</strong>d water tr<strong>an</strong>sport. The relatively high<br />

porosity <strong>of</strong> the soil structure (Plate 6.31) could therefore be partly attributed to binding <strong>an</strong>d<br />

cementation <strong>of</strong> clay matrix by iron oxide to form detached lumps; <strong>an</strong>d partly to leaching <strong>of</strong><br />

soluble bases <strong>of</strong> Ca, Mg, K, Na; as well as solution effects caused by Fe going into reaction.<br />

Plate 6.25. Loose, porous <strong>an</strong>d friable structure<br />

<strong>of</strong> red soils.<br />

Plate 6.26. Lumps <strong>of</strong> cemented clay<br />

matrix with Fe-rich crusts.<br />

The fine matrix shows presence <strong>of</strong> Si <strong>an</strong>d Al mainly, suggesting kaolinite; as well as Fe,<br />

implying heamatite. Fine crusts bearing Ti, Mn <strong>an</strong>d K also occur (Plate 6.27) . Kaolinite<br />

generally occurs as fine weathered <strong>an</strong>d altered plates which have lost the original six-sided<br />

form <strong>an</strong>d/ or characteristic crystal structure. The plates are commonly weakly bound<br />

/cemented together <strong>an</strong>d covered by fine Fe-rich crusts <strong>of</strong> matrix (Plate 6.28). Weathering <strong>an</strong>d


76<br />

alteration <strong>of</strong> kaolinite is most likely caused by low pH conditions that result in Aluminium<br />

depletion (Table 6.1). Remains <strong>of</strong> org<strong>an</strong>ic matter in the form <strong>of</strong> pl<strong>an</strong>t roots are also found<br />

(Plate 6.27). On the whole, presence <strong>of</strong> quartz is limited, while that <strong>of</strong> feldspar is not evident<br />

<strong>an</strong>d this could be attributed to possible weathering <strong>an</strong>d alteration.<br />

Plate 6.27. Org<strong>an</strong>ic matter intermingled with Fe-rich lumps <strong>of</strong> clay matrix<br />

(matrix composition in diagram).<br />

Plate 6.28. Abraded sub-rounded to rounded quartz grain in clay matrix <strong>of</strong> obscure fine<br />

weathered kaolinite plates covered with Fe-rich crust (composition <strong>of</strong> matrix in diagram).


77<br />

Plate 6.29. Iron concretion with elevated Fe content.<br />

Plate 6.30. Solution pores/ cavities in<br />

weathered iron concretion.<br />

Plate 6.31. Porous structure <strong>of</strong> red soils.<br />

6.5 Import<strong>an</strong>ce <strong>an</strong>d influence <strong>of</strong> clay minerals on <strong>engineering</strong> properties <strong>of</strong> black <strong>clays</strong><br />

<strong>an</strong>d red soils<br />

Clay minerals are usually platy with a surface negative charge (Johnson <strong>an</strong>d Degraff, 1988).<br />

In the presence <strong>of</strong> water, electrical attraction <strong>of</strong> opposing charges on water molecules causes<br />

the clay particles to form clusters <strong>an</strong>d/ or flocculate <strong>an</strong>d settle down in the suspension.<br />

According to Day (2001), clay minerals serve to influence soil <strong>engineering</strong> properties such as<br />

plasticity, swelling, shrinkage, shear strength, consolidation <strong>an</strong>d permeability. They therefore<br />

account for the increase in volume <strong>an</strong>d/ or swelling which occurs when clay soils are allowed<br />

free access to water. Different clay minerals exhibit stronger or weaker negative charges <strong>an</strong>d<br />

this in turn determine the relative degree <strong>an</strong>d ease with which they flocculate <strong>an</strong>d cause the<br />

clay soils involved to swell/ exp<strong>an</strong>d, i.e. the reactivity <strong>of</strong> the clay soil.<br />

Clay mineralogy is <strong>an</strong> import<strong>an</strong>t controlling factor <strong>of</strong> the plasticity <strong>of</strong> clay soils (Johnson <strong>an</strong>d<br />

Degraff, 1988). In practice, kaolinite <strong>an</strong>d illite are classified as nonexp<strong>an</strong>sive clay minerals<br />

<strong>an</strong>d are characterised by relatively lower values <strong>of</strong> plasticity indices (PI). According to Holtz<br />

<strong>an</strong>d Kovacs (1981), kaolinite is a relatively inactive clay mineral so that even though it is


78<br />

technically a clay, it behaves more as a silt material. In this study, the red soils were found to<br />

show relatively lower plasticity indices (18-22%) as well as lower swelling capabilities (free<br />

swell: 15-20%) <strong>an</strong>d activity values (0,83-1,03); <strong>an</strong>d this could be attributed to the higher<br />

kaolinite content (80% <strong>an</strong>d over) which characterises these soils.<br />

On the other h<strong>an</strong>d, montmorillonite belongs to a group <strong>of</strong> clay minerals that are characterised<br />

by weakly bonded layers (Day, 2001). Each layer consists <strong>of</strong> two silica sheets with <strong>an</strong><br />

aluminium (gibbsite) sheet in the middle. Water <strong>an</strong>d exch<strong>an</strong>geable cations (Na, Ca, etc.) c<strong>an</strong><br />

enter <strong>an</strong>d separate the layers, creating a very small crystal that has a strong attraction for<br />

water. As a result, montmorillonite has the highest activity (A = 4-7) <strong>an</strong>d usually exhibits the<br />

highest water content, greatest compressibility, <strong>an</strong>d lowest shear strength <strong>of</strong> all the clay<br />

minerals.<br />

Montmorillonite usually harbours a relatively larger negative surface charge <strong>an</strong>d readily<br />

attracts water molecules by adsorption (Johnson <strong>an</strong>d Degraff, 1988). It is therefore classified<br />

as exp<strong>an</strong>sive <strong>an</strong>d would usually exhibit a wide r<strong>an</strong>ge <strong>of</strong> PI values depending on adsorbed<br />

cations available to satisfy its charge deficiencies. The charge deficiencies are better satisfied<br />

by calcium-rich pore water (due to higher calcium valence <strong>of</strong> 2) which also results in<br />

improved interparticle bonding so that the amount <strong>of</strong> water needed for complete charge<br />

neutralisation would be considerably reduced. As a result, calcium-rich montmorillonite<br />

would exhibit relatively lower plasticity indices. On the other h<strong>an</strong>d, sodium-rich pore water<br />

would result in sodium-rich montmorillonite which usually requires more water (due to lower<br />

valence <strong>of</strong> 1) for complete charge neutralisation so that the plasticity indices are also<br />

comparatively higher. The black <strong>clays</strong> involved in this study are characterised by<br />

comparatively higher plasticity indices (39-55%), swelling capabilities (free swell: 100-<br />

145%) <strong>an</strong>d activity values (1,10-3,38); <strong>an</strong>d this could be related to the higher content <strong>of</strong><br />

smectites (90% <strong>an</strong>d over) in these soils. Results <strong>of</strong> mineralogical <strong>an</strong>alyses obtained in this<br />

study suggest the black <strong>clays</strong> to be more <strong>of</strong> calcium-rich montmorillonite th<strong>an</strong> sodium-rich<br />

montmorillonite.<br />

Illite has a structure similar to that <strong>of</strong> montmorillonite, but the layers are more strongly<br />

bonded together (Day, 2001; Holtz & Kovacs, 1981). It is usually intermediate in activity (A<br />

= 0,5-1,3) between kaolinite <strong>an</strong>d montmorillonite in terms <strong>of</strong> its cation exch<strong>an</strong>ge capacity,<br />

ability to absorb <strong>an</strong>d retain water, as well as physical characteristics such as plasticity index;<br />

<strong>an</strong>d <strong>of</strong>ten plots just above the A-line on the plasticity chart. In this study, traces <strong>of</strong> illite were<br />

noted in black <strong>clays</strong>, <strong>an</strong>d are most probably a result <strong>of</strong> alteration <strong>of</strong> feldspars <strong>an</strong>d / or<br />

kaolinite.<br />

6.6 Carbon <strong>an</strong>d sulphur content <strong>of</strong> soils<br />

6.6.1 Method <strong>of</strong> <strong>an</strong>alysis<br />

Determination <strong>of</strong> carbon <strong>an</strong>d sulphur contents <strong>of</strong> soil was carried out using the Leco CS-225<br />

apparatus under conditions <strong>of</strong> pressurised gas (99,5% oxygen). Two st<strong>an</strong>dards were employed<br />

in the <strong>an</strong>alysis, i.e.<br />

Leco-Iron filings 3,32%C + 0,063%S, <strong>an</strong>d<br />

Leco-Steel-ring specimen 0,358%C + 0,0178%S.


79<br />

The CS-225 is a micro-processor <strong>an</strong>alysis system for determination <strong>of</strong> carbon <strong>an</strong>d sulphur<br />

contents through infrared absorption. Iron <strong>an</strong>d tungsten filings acting as catalysts were added<br />

to the soil specimen which was then heated in a high frequency induction oven. The<br />

combustion/ heating gases were dried <strong>an</strong>d then directed to the infrared sulphur cell for the<br />

measurement <strong>of</strong> the sulphur dioxide content. The gases were further passed through a<br />

catalysing oven where carbon monoxide (CO) <strong>an</strong>d sulphur dioxide (SO2) were converted into<br />

carbon dioxide (CO2) <strong>an</strong>d sulphur trioxide SO3, respectively. The SO3 was collected in a<br />

sulphur-trap, thereby permitting measurement <strong>of</strong> the total carbon content in terms <strong>of</strong> CO2 in<br />

the infrared carbon cell. The results <strong>of</strong> the <strong>an</strong>alysis were calculated by the microprocessor<br />

system with the aid <strong>of</strong> calibration factors <strong>an</strong>d after applying relev<strong>an</strong>t weight compesations.<br />

The org<strong>an</strong>ic carbon content was determined by first pretreating the soil specimen with<br />

concentrated hydrochloric acid to remove the carbonate part. The excess acid was evaporated<br />

by heating on a hot plate <strong>an</strong>d finally removed by passing over potassium iodide <strong>an</strong>d <strong>an</strong>timony<br />

metal.<br />

The sulphur content <strong>of</strong> both the red soils <strong>an</strong>d black <strong>clays</strong> was found to be negligible. The<br />

results are therefore not reported.<br />

6.6.2 Results<br />

Results <strong>of</strong> carbon content determinations <strong>of</strong> soils performed in this study are given in Table<br />

6.6. The black <strong>clays</strong> exhibit 0,50-1,48% total carbon (me<strong>an</strong>: 1,17%) <strong>an</strong>d 0,22-1,19% org<strong>an</strong>ic<br />

carbon (me<strong>an</strong>: 0,93%); <strong>an</strong>d therefore 0,04-0,74% inorg<strong>an</strong>ic carbon (me<strong>an</strong>: 0,24%). The red<br />

soils have 0,45-2,47% total carbon (me<strong>an</strong>: 1,12%), 0,35-1,90 org<strong>an</strong>ic carbon (me<strong>an</strong>: 0,85%)<br />

<strong>an</strong>d 0,10-0,57% inorg<strong>an</strong>ic carbon (me<strong>an</strong>: 0,27%). The total carbon content in both types <strong>of</strong><br />

soil is generally low (less th<strong>an</strong> 2,5%) so that difficulties were encountered in trying to<br />

separate <strong>an</strong>d determine the org<strong>an</strong>ic <strong>an</strong>d inorg<strong>an</strong>ic components during laboratory <strong>an</strong>alysis. In<br />

this study, therefore, the total carbon <strong>an</strong>d org<strong>an</strong>ic carbon contents were determined; <strong>an</strong>d their<br />

numerical difference adopted as the inorg<strong>an</strong>ic carbon content, i.e.<br />

Total carbon (%) = org<strong>an</strong>ic carbon (%) + inorg<strong>an</strong>ic carbon (%), or<br />

Ctotal (%) = Corg<strong>an</strong>ic (%) + Cdiff. (%),<br />

where the numerical difference, Cdiff., corresponds to the inorg<strong>an</strong>ic carbon component.<br />

The org<strong>an</strong>ic carbon component serves to represent the org<strong>an</strong>ic matter content <strong>of</strong> the soils. The<br />

inorg<strong>an</strong>ic carbon component represents the carbonate content occurring in the soils in the<br />

form <strong>of</strong> calcite, dolomite, <strong>an</strong>kerite, siderite <strong>an</strong>d/or other inorg<strong>an</strong>ic carbonates. The Scheibler<br />

method for determination <strong>of</strong> carbonate content <strong>of</strong> soils could not be applied in the present<br />

<strong>an</strong>alysis due to the generally low carbonate content (less th<strong>an</strong> 5%) <strong>of</strong> the black <strong>clays</strong> <strong>an</strong>d red<br />

soils. This <strong>an</strong>alysis method depends on atmospheric conditions <strong>of</strong> temperature <strong>an</strong>d pressure,<br />

<strong>an</strong>d is therefore only applicable to soils with carbonate contents <strong>of</strong> over 5%.


80<br />

Table 6.6. Carbon content <strong>of</strong> soils <strong>of</strong> the study area.<br />

Sample No. Total carbon;<br />

Ctotal (%)<br />

Org<strong>an</strong>ic carbon;<br />

Corg<strong>an</strong>ic (%)<br />

Inorg<strong>an</strong>ic carbon;<br />

Cdiff. (%)<br />

SA1-30cm 1,16 1,04 0,12<br />

SA1-50cm 1,48 0,81 0,67<br />

SA2-70cm 0,50 0,46 0,04<br />

Black <strong>clays</strong> SA2-105cm 0,96 0,22 0,74<br />

SA41-50cm 1,14 1,02 0,12<br />

SB1-50cm 1,13 1,00 0,13<br />

SB1-70cm 1,19 1,08 0,11<br />

SB42-30cm 1,34 1,19 0,15<br />

SB42-50cm 1,30 1,17 0,13<br />

SC17-50cm 1,29 1,08 0,21<br />

SC33-30cm 1,39 1,19 0,20<br />

Rd1-30cm 2,47 1,90 0,57<br />

Red soils Rd1-100cm 0,73 0,54 0,19<br />

Rd1-200cm 0,83 0,61 0,22<br />

Rd1-400cm 0,45 0,35 0,10<br />

6.6.3 Analysis <strong>of</strong> results<br />

The variation <strong>of</strong> carbon content <strong>of</strong> soils with depth, Dp, is represented diagrammatically in<br />

Figures 6.11, 6.12 <strong>an</strong>d 6.13.<br />

Red soils: carbon content/ depth<br />

3,00<br />

carbon (%)<br />

2,50<br />

2,00<br />

1,50<br />

1,00<br />

0,50<br />

Ctotal = 1,0356Dp -0,6115<br />

R 2 = 0,8831<br />

Corg = 0,7821Dp -0,6116<br />

R 2 = 0,8766<br />

total carbon<br />

org<strong>an</strong>ic carbon<br />

inorg<strong>an</strong>ic carbon<br />

Potential (total carbon)<br />

Potential (org<strong>an</strong>ic<br />

carbon)<br />

Potential (inorg<strong>an</strong>ic<br />

carbon)<br />

Cdiff = 0,2526Dp -0,6128<br />

0,00<br />

R 2 = 0,8871<br />

0,00 1,00 2,00 3,00 4,00 5,00<br />

depth Dp (m)<br />

Figure 6.11. Variation <strong>of</strong> carbon content <strong>of</strong> red soils with depth.


81<br />

The red soils generally exhibit a decrease in their carbon contents with depth (Fig. 6.11). The<br />

decrease <strong>of</strong> the org<strong>an</strong>ic carbon is indicative <strong>of</strong> a general decrease <strong>of</strong> the org<strong>an</strong>ic matter content<br />

<strong>of</strong> the soils with depth. The decrease <strong>of</strong> the inorg<strong>an</strong>ic carbon could be related to the good<br />

drainage conditions <strong>of</strong> the soils which favour leaching <strong>an</strong>d removal <strong>of</strong> carbonates as well as<br />

other soluble components <strong>of</strong> Mg, Ca, Na <strong>an</strong>d K. The variation <strong>of</strong> total carbon, org<strong>an</strong>ic carbon<br />

<strong>an</strong>d inorg<strong>an</strong>ic carbon contents <strong>of</strong> these soils with depth is best described by a potential<br />

relationship with strong correlations (R = 0,94; R = 0,94; <strong>an</strong>d R=0,94, respectively).<br />

Generally, however, the soils exhibit higher contents <strong>of</strong> org<strong>an</strong>ic carbon th<strong>an</strong> inorg<strong>an</strong>ic carbon.<br />

The total carbon <strong>an</strong>d org<strong>an</strong>ic carbon contents <strong>of</strong> black <strong>clays</strong> generally decrease with depth<br />

(Fig. 6.12); <strong>an</strong>d this is most probably due to the decrease <strong>of</strong> org<strong>an</strong>ic matter content <strong>of</strong> the soils<br />

with depth. On the contrary, the inorg<strong>an</strong>ic carbon component is found to generally increase<br />

with depth, <strong>an</strong>d this could be attributed to the weathering <strong>of</strong> the underlying volc<strong>an</strong>ic tuffs<br />

which usually results in the formation <strong>of</strong> secondary limestone in situ. The variation <strong>of</strong> total<br />

carbon with depth could be described by a polynomial relationship, but with a relatively<br />

weaker correlation (R = 0,56); while the variation <strong>of</strong> org<strong>an</strong>ic carbon with depth is best<br />

approximated by <strong>an</strong> exponetial relationship with a strong correlation (R = 0,85). On the other<br />

h<strong>an</strong>d, variation <strong>of</strong> inorg<strong>an</strong>ic carbon content with depth tends to fit a polynomial relationship,<br />

with a moderately strong correlation (R = 0,63).<br />

Black <strong>clays</strong>: carbon content/ depth<br />

1,60<br />

1,40<br />

1,20<br />

Ctotal = 0,5235Dp 2 - 1,3301Dp + 1,7069<br />

total carbon<br />

carbon (%)<br />

1,00<br />

0,80<br />

0,60<br />

0,40<br />

R 2 = 0,3146<br />

Corg = 2,4918e -2,0216Dp<br />

R 2 = 0,7303<br />

CDiff = 1,5224Dp 2 - 1,4352Dp + 0,5028<br />

org<strong>an</strong>ic carbon<br />

inorg<strong>an</strong>ic carbon<br />

Polynomial (total<br />

carbon)<br />

Exponential<br />

(org<strong>an</strong>ic carbon)<br />

Polynomial<br />

(inorg<strong>an</strong>ic carbon)<br />

0,20<br />

R 2 = 0,3943<br />

0,00<br />

0,00 0,20 0,40 0,60 0,80 1,00 1,20<br />

depth Dp (m)<br />

Figure 6.12. Variation <strong>of</strong> carbon content <strong>of</strong> black <strong>clays</strong> with depth.<br />

Combined results <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils also show a general decrease <strong>of</strong> total carbon as<br />

well as org<strong>an</strong>ic carbon with depth (Fig. 6.13). The results also reveal approximating potential<br />

relationships with strong correlations in the variation <strong>of</strong> total carbon (R = 0,77) <strong>an</strong>d org<strong>an</strong>ic<br />

carbon (R = 0,75) with depth. However, no well defined relationship occurs to describe the<br />

variation <strong>of</strong> combined data <strong>of</strong> inorg<strong>an</strong>ic carbon <strong>of</strong> the two types <strong>of</strong> soil with depth.


82<br />

Combined black <strong>an</strong>d red soils: carbon content/ depth<br />

3,00<br />

2,50<br />

2,00<br />

carbon (%)<br />

1,50<br />

Ctotal = 0,8709Dp -0,4442<br />

R 2 = 0,5916<br />

total carbon<br />

org<strong>an</strong>ic carbon<br />

inorg<strong>an</strong>ic carbon<br />

Potential (total carbon)<br />

Potential (org<strong>an</strong>ic carbon)<br />

1,00<br />

0,50<br />

Corg = 0,618Dp -0,5706<br />

R 2 = 0,5645<br />

0,00<br />

0,00 1,00 2,00 3,00 4,00 5,00<br />

depth Dp (m)<br />

Figure 6.13. Variation <strong>of</strong> carbon content <strong>of</strong> both black <strong>clays</strong> <strong>an</strong>d red soils with depth.


83<br />

Chapter 7<br />

Laboratory soils index <strong>an</strong>d <strong>engineering</strong> properties determination<br />

7.1 Moisture content <strong>an</strong>d index tests<br />

7.1.1 Introduction<br />

The perform<strong>an</strong>ce <strong>of</strong> soil index tests usually involves the measurement <strong>of</strong> moisture content <strong>of</strong><br />

soils, both in the natural state as well as under certain defined test conditions (i.e. consistency<br />

limits). Index tests serve to investigate the way in which the amount <strong>of</strong> water in soils c<strong>an</strong><br />

influence their behaviour. They also provide a useful method <strong>of</strong> classifying cohesive soils <strong>an</strong>d<br />

<strong>of</strong> assessing their <strong>engineering</strong> properties (Head, 1984).<br />

The index properties <strong>of</strong> soils investigated in this study include natural moisture content,<br />

Atterberg limits (liquid <strong>an</strong>d plastic limits, plasticity index), swelling capability (free swell)<br />

<strong>an</strong>d linear shrinkage.<br />

Natural moisture content <strong>an</strong>d Atterberg limits usually serve to describe <strong>an</strong>d provide a<br />

sufficient underst<strong>an</strong>ding <strong>of</strong> the nature <strong>of</strong> clay soils, as may be required by m<strong>an</strong>y applications<br />

in <strong>engineering</strong> practice; <strong>an</strong>d this especially when coupled with a knowledge <strong>of</strong> the <strong>geological</strong><br />

history <strong>of</strong> the soils (Atterberg, 1911; Nelson <strong>an</strong>d Miller, 1992).<br />

7.1.2 Natural moisture content<br />

Determination <strong>of</strong> natural moisture content <strong>of</strong> soils was carried out in this study by ovendrying<br />

method <strong>an</strong>d in accord<strong>an</strong>ce with British St<strong>an</strong>dard (BS 1377; 1975, Test 1(A)) <strong>an</strong>d<br />

Germ<strong>an</strong> St<strong>an</strong>dard, i.e. Deutsches Institut für Normung (DIN 18 121, Teil 1). This testing<br />

procedure specifies a st<strong>an</strong>dard drying temperature <strong>of</strong> 105-110 °C.<br />

Soil samples used for natural moisture content determination had been previously collected<br />

from the field, whereupon they were immediately well sealed in polythene bags to prevent<br />

moisture loss. The moisture content so determined also represents the moisture content <strong>of</strong><br />

natural undisturbed soil in situ.<br />

Natural moisture content, Wn, <strong>of</strong> a soil is usually expressed as a percentage <strong>of</strong> its dry mass<br />

(Head, 1984); <strong>an</strong>d is given by the equation<br />

natural moisture content = (loss <strong>of</strong> moisture/dry mass) * 100%, or<br />

Wn (%) = ((m2-m3)/(m3-m1)) * 100 (7.1)<br />

where m1 = mass <strong>of</strong> empty container<br />

m2 = mass <strong>of</strong> container + wet soil<br />

m3 = mass <strong>of</strong> container + dry soil<br />

In this study, up to three separate moisture content determinations were performed on each<br />

soil sample <strong>an</strong>d the average value calculated. Results <strong>of</strong> natural moisture content<br />

determination <strong>of</strong> soils performed in this study are presented in Table (7.2); <strong>an</strong>d are reported to<br />

the nearest whole number or 1%.


84<br />

7.1.3 Atterberg limits <strong>an</strong>d consistency <strong>of</strong> clay soils<br />

7.1.3.1 Scope<br />

The Atterberg limits serve to define the r<strong>an</strong>ge <strong>of</strong> moisture contents within which a clay soil<br />

exhibits a plastic consistency by measuring <strong>an</strong>d describing the plasticity r<strong>an</strong>ge in numerical<br />

terms (Nelson <strong>an</strong>d Miller, 1992).<br />

The Atterberg limits <strong>of</strong> measurement adopted in this study to describe the consistency <strong>of</strong> soils<br />

include the liquid limit (LL) <strong>an</strong>d the plastic limit (PL). The liquid limit is the moisture content<br />

at which a soil passes from the plastic to the liquid state, while the plastic limit represents the<br />

moisture content at which a soil passes from the plastic to the solid state <strong>an</strong>d becomes too dry<br />

to be in a plastic condition. The plasticity index (PI) is the moisture content r<strong>an</strong>ge between the<br />

plastic <strong>an</strong>d liquid limits, <strong>an</strong>d serves as a measure <strong>of</strong> the plasticity <strong>of</strong> the clay soil (Nelson <strong>an</strong>d<br />

Miller, 1992).<br />

In this study, the state <strong>of</strong> consistency <strong>of</strong> a clay soil in its natural state was numerically<br />

described by relating its natural moisture content to the plastic <strong>an</strong>d liquid limits using two<br />

parameters, i.e. the relative consistency (Cr) <strong>an</strong>d the liquidity index (LI). This is because in<br />

general, clay soils occur as a slurry (liquid state) when moisture contents exceed their liquid<br />

limits, while they exhibit a firm <strong>an</strong>d plastic consistency with moisture contents lying within<br />

the plasticity r<strong>an</strong>ge, i.e. between plastic <strong>an</strong>d liquid limits. On the other h<strong>an</strong>d, clay soils with<br />

moisture contents below their plastic limits would usually exhibit a hard to stiff consistency<br />

(Head, 1984). According to Terzaghi <strong>an</strong>d Peck (1948), relative consistency, Cr, is the ratio <strong>of</strong><br />

the difference between liquid limit <strong>an</strong>d moisture content to the plasticity index, i.e.<br />

Cr = (LL-W)/(LL-PL), or<br />

Cr = (LL-W)/PI (7.2)<br />

where W = moisture content<br />

Lambe <strong>an</strong>d Whitm<strong>an</strong> (1969) defined the liquidity index, LI, as the ratio <strong>of</strong> the difference<br />

between moisture content <strong>an</strong>d plastic limit to the plasticity index, i.e.<br />

LI = W-PL/PI (7.3)<br />

According to Skempton <strong>an</strong>d Northey (1953) <strong>an</strong>d Wood <strong>an</strong>d Wroth (1978), clay soils usually<br />

exhibit increasing shear strength values as their state <strong>of</strong> consistency varies with progressive<br />

decrease <strong>of</strong> moisture content, starting from the liquid limit down to the plastic limit. This<br />

study tries to investigate <strong>an</strong>d establish <strong>an</strong>y possible consistent form <strong>of</strong> relationship between<br />

shear strength parameters <strong>an</strong>d liquidity index <strong>an</strong>d/ or relative consistency for the clay soils<br />

found.<br />

Skempton (1953) showed that the Atterberg limits <strong>of</strong> <strong>clays</strong> are related to the particle size <strong>an</strong>d<br />

clay mineralogical composition <strong>of</strong> the soils. He noted that the the plasticity index, PI, is<br />

dependent on the proportion <strong>of</strong> clay fraction, <strong>an</strong>d defined colloidal activity (A) <strong>of</strong> a clay soil<br />

as the ratio between these two parameters, i.e.<br />

activity A = PI/ clay fraction (7.4)


85<br />

According to Johnson <strong>an</strong>d Degraff (1988), clay minerals are usually platy with a negative<br />

surface charge which causes them to readily adsorp water. The presence <strong>of</strong> clay minerals in a<br />

soil therefore causes it to increase in volume or exp<strong>an</strong>d in the presence <strong>of</strong> water. Different<br />

clay minerals exhibit stronger or weaker negative charges which in turn serve as a measure <strong>of</strong><br />

the activity <strong>of</strong> the clay soil in terms <strong>of</strong> the relative degree with which the negative charge<br />

would influence opposing charges in adsorped water to cause the clay particles to form<br />

clusters or flocculate <strong>an</strong>d settle down in a suspension <strong>of</strong> water. As a result, the amount <strong>an</strong>d<br />

type <strong>of</strong> clay minerals present in a soil have a signific<strong>an</strong>t effect on soil <strong>engineering</strong> properties<br />

such as plasticity, swelling, shrinkage, shear strength, consolidation <strong>an</strong>d permeability.<br />

Table 7.1. Activity <strong>of</strong> clay soils <strong>an</strong>d clay minerals (After Skempton, 1953;<br />

Mitchell, 1976; <strong>an</strong>d Day, 2001).<br />

CLAY SOILS<br />

ACTIVITY<br />

Inactive <strong>clays</strong> < 0,75<br />

Normal <strong>clays</strong> 0,75-1,25<br />

Active <strong>clays</strong> 1,25-2,0<br />

Highly active <strong>clays</strong> > 2,0<br />

Extremely active <strong>clays</strong>, e.g. bentonite 6 or more<br />

CLAY MINERALS<br />

APPROX. ACTIVITY<br />

Kaolinite 0,3-0,5<br />

Il lite 0,5-1,3<br />

Na-montmorillonite 4,0-7,0<br />

Ca-montmorillonite 1,5<br />

Gr<strong>an</strong>ular soils, quartz 0<br />

THIS STUDY<br />

ACTIVITY<br />

Black <strong>clays</strong> 1,1 – 3,38 (me<strong>an</strong> value: 1,87)<br />

Red soils 0,83 – 1,03 (me<strong>an</strong> value: 0,94)<br />

The colloidal activity is usually more or less const<strong>an</strong>t for a particular type <strong>of</strong> clay soil. A<br />

suggested classification <strong>of</strong> clay soils based on colloidal activity is given in Table (7.1). A<br />

similar classification <strong>of</strong> common clay minerals is also included in the table. R<strong>an</strong>ges <strong>of</strong> activity<br />

values for black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study are also given for comparison purposes. A<br />

whole list <strong>of</strong> calculated activity values from results <strong>of</strong> plasticity indices <strong>an</strong>d grain size <strong>an</strong>alysis<br />

obtained in this study is included in Table (7.3). An activity classification <strong>of</strong> the soils in this<br />

study is also presented alongside. The activity state <strong>an</strong>d/ or activity level <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d<br />

red soils <strong>of</strong> this study is illustrated diagrammatically in the activity chart (Fig. 7.2, after<br />

NAVFAC, 1982).


86<br />

7.1.3.2 Plastic limit<br />

Plastic limit tests were carried out in this study to determine the lowest moisture content at<br />

which the clay soils found are plastic. The tests were carried out according to British St<strong>an</strong>dard<br />

(BS 1377: 1975, Test 3) <strong>an</strong>d Germ<strong>an</strong> St<strong>an</strong>dard (DIN 18 122, Teil 1); by using a suitable<br />

amount <strong>of</strong> disturbed soil previously oven-dried (105 °C over a 12-24 hour period) <strong>an</strong>d sieved<br />

to provide a homogeneous material.<br />

The testing method is shown in Plate (7.1) in which a uniform pressure was used to roll<br />

threads <strong>of</strong> a homogeneous soil paste between the fingers <strong>of</strong> one h<strong>an</strong>d <strong>an</strong>d a glass plate from<br />

about 6 mm diameter size until they just crumbled at about 3 mm diameter. The moisture<br />

content <strong>of</strong> the threads at the first crumbling point was determined. In this study, a total <strong>of</strong><br />

three separate moisture content determinations <strong>of</strong> the threads was carried out for each soil<br />

sample <strong>an</strong>d <strong>an</strong> average value calculated <strong>an</strong>d taken as the plastic limit (PL) <strong>of</strong> the soil. Results<br />

<strong>of</strong> plastic limit <strong>of</strong> soils obtained in this study are reported to the nearest whole number (1%)<br />

alongside those <strong>of</strong> liquid limit <strong>an</strong>d plasticity index in Table (7.2).<br />

Plate 7.1. Plastic limit determination in this study.<br />

7.1.3.3 Liquid limit<br />

Liquid limit determination <strong>of</strong> soils in this study was carried out using the Casagr<strong>an</strong>de<br />

apparatus (Casagr<strong>an</strong>de, 1958), <strong>an</strong>d in accord<strong>an</strong>ce with British St<strong>an</strong>dard (BS 1377: 1975, Test<br />

2(B)) <strong>an</strong>d Germ<strong>an</strong> St<strong>an</strong>dard (DIN 18 122, Teil 1). The apparatus used in this study is shown<br />

in Plate (7.2). The test involved applying bumps or blows to the cup <strong>of</strong> the apparatus so as to<br />

close a V-shaped groove cut in the middle <strong>of</strong> a homogeneous soil paste contained in the cup.<br />

The moisture content <strong>of</strong> the paste at which a total <strong>of</strong> 25 bumps closed the groove along a<br />

continuous dist<strong>an</strong>ce <strong>of</strong> 10 mm was determined <strong>an</strong>d recorded as the liquid limit (LL) <strong>of</strong> the<br />

soil.


87<br />

Table 7.2. Results <strong>of</strong> index tests performed on black <strong>clays</strong> <strong>an</strong>d red soils in this study.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%) PI (%) Cr LI FS (%) LS (%)<br />

PI/LL class<br />

(BS 1377) Consistency<br />

SA1-30cm 31,6 78,2 37,2 41,1 1,136 -0,14 135 25 MV semi-solid<br />

A SA1-50cm 30 74,9 36,3 38,6 1,163 -0,16 100 23 MV semi-solid<br />

SA2-30cm 33,7 92,6 38,8 53,9 1,094 -0,09 135 27 CE semi-solid<br />

(black <strong>clays</strong>) SA2-70cm 34,1 90 36,3 53,7 1,04 -0,04 170 25 CV-CE semi-solid<br />

SA2-105cm 29,7 88,1 32,9 55,2 1,059 -0,06 120 26 CV semi-solid<br />

SA28-30cm 25,6 84,8 39,7 45,1 1,313 -0,31 125 25 MV-CV semi-solid<br />

SA28-50cm 26,8 88,5 41,3 47,2 1,307 -0,31 135 25 MV-CV semi-solid<br />

SA37-30cm 19,4 80,9 38,1 42,7 1,439 -0,44 115 24 CV semi-solid<br />

SA37-50cm 21,8 83,3 39,2 44,2 1,393 -0,39 115 26 CV semi-solid<br />

SA37-80cm 23,1 85,1 39,2 45,9 1,35 -0,35 115 26 CV semi-solid<br />

SA41-30cm 23 88,5 41,3 47,2 1,387 -0,39 135 26 CV semi-solid<br />

SA41-50cm 24,9 89,7 41,1 48,6 1,334 -0,33 120 24 CV semi-solid<br />

SB1-30cm 23 81,7 34,3 47,4 1,237 -0,24 125 23 CV semi-solid<br />

B SB1-50cm 26,7 83,7 36,1 47,6 1,197 -0,2 135 24 CV semi-solid<br />

SB1-70cm 29,7 89,9 39,1 50,8 1,187 -0,19 145 25 CV semi-solid<br />

(black <strong>clays</strong>) SB5-30cm 24 80,9 34 47 1,212 -0,21 120 23 CV semi-solid<br />

SB5-50cm 26,8 84,6 36,6 48 1,205 -0,2 135 24 CV semi-solid<br />

SB5-70cm 30,5 91,1 40,2 51 1,19 -0,19 145 26 CE semi-solid<br />

SB7-30cm 23,2 82,9 34,6 48,3 1,235 -0,23 130 24 CV semi-solid<br />

SB7-50cm 26,1 85,8 37 48,8 1,222 -0,22 135 24 CV semi-solid<br />

SB7-70cm 29,6 87 38,3 48,7 1,179 -0,18 140 25 CV semi-solid<br />

SB18-30cm 20 82,4 39,8 42,6 1,463 -0,46 120 23 MV-CV semi-solid<br />

SB18-50cm 20,1 72,4 33,6 38,8 1,346 -0,35 120 21 CV semi-solid<br />

SB27-30cm 21,6 85 38 47,1 1,348 -0,35 130 24 CV semi-solid<br />

SB27-50cm 22,2 87,7 39,2 48,5 1,352 -0,35 130 25 CV semi-solid<br />

SB31-30cm 23,4 86,9 38,1 48,8 1,3 -0,3 130 24 CV semi-solid<br />

SB31-50cm 22,9 86,1 37,6 48,4 1,304 -0,3 130 25 CV semi-solid<br />

SB41-30cm 23 88,8 44,6 44,1 1,491 -0,49 145 26 MV semi-solid<br />

SB41-50cm 24 90,7 43,6 47,1 1,415 -0,42 145 25 MV semi-solid<br />

SB42-30cm 22 87,6 43,8 43,8 1,499 -0,5 145 25 MV semi-solid<br />

SB42-50cm 21,3 91,1 44,5 46,6 1,498 -0,5 145 26 ME-CE semi-solid<br />

SC1-30cm 24,2 82 34,5 47,5 1,218 -0,22 125 23 CV semi-solid<br />

C SC1-50cm 25,9 84,9 36,7 48,2 1,224 -0,22 135 24 CV semi-solid<br />

SC5-30cm 23 83,1 34,7 48,4 1,24 -0,24 125 23 CV semi-solid<br />

(black <strong>clays</strong>) SC5-50cm 25,1 83,5 36,2 47,2 1,236 -0,24 135 24 CV semi-solid<br />

SC9-30cm 24,8 83,6 35,6 48 1,223 -0,22 130 24 CV semi-solid<br />

SC9-50cm 26 84,6 37,1 47,5 1,233 -0,23 135 24 CV semi-solid<br />

SC17-30cm 21,2 75,7 34,7 41 1,329 -0,33 105 25 CV semi-solid<br />

SC17-50cm 22 80,7 32,5 48,2 1,219 -0,22 110 24 CV semi-solid<br />

SC25-30cm 26,1 87,9 39 48,9 1,265 -0,27 125 25 CV semi-solid<br />

SC25-50cm 25,8 88,6 41 47,6 1,321 -0,32 145 25 MV semi-solid<br />

SC29-30cm 20,7 81,5 32,2 49,3 1,235 -0,23 110 25 CV semi-solid<br />

SC29-50cm 20,4 80,2 36,8 43,4 1,379 -0,38 110 26 CV semi-solid<br />

SC33-30cm 20,3 81,7 32,5 49,1 1,248 -0,25 120 25 CV semi-solid<br />

SC33-50cm 21,4 82,9 33 49,9 1,233 -0,23 120 26 CV semi-solid<br />

SC41-30cm 17,7 85,2 37,8 47,4 1,422 -0,42 125 29 CV semi-solid<br />

SC41-50cm 17,2 83,6 36,6 47 1,412 -0,41 125 29 CV semi-solid


88<br />

Table 7.2, continued. Results <strong>of</strong> index tests performed on black <strong>clays</strong> <strong>an</strong>d red soils in this<br />

study.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%) PI (%) Cr LI FS (%) LS (%)<br />

PI/LL class<br />

(BS 1377) Consistency<br />

SD1-30cm 26,5 82,3 35,9 46,4 1,203 -0,2 125 26 CV semi-solid<br />

D SD1-50cm 27,1 83 36,1 46,9 1,19 -0,19 130 27 MV-CV semi-solid<br />

SD1-80cm 29,5 91,1 41,7 49,4 1,248 -0,25 140 26 ME-CE semi-solid<br />

(black <strong>clays</strong>) SD5-30cm 26,2 83,5 36,4 47,1 1,217 -0,22 125 26 CV semi-solid<br />

SD5-50cm 27,4 84,2 36,4 47,9 1,188 -0,19 130 27 CV semi-solid<br />

SD5-80cm 29,7 92,3 42 50,4 1,243 -0,24 140 27 ME-CE semi-solid<br />

SD7-30cm 26,9 84,3 36,4 48 1,198 -0,2 110 27 CV semi-solid<br />

SD7-50cm 27,1 85,9 37,1 48,8 1,205 -0,2 140 26 CV semi-solid<br />

SD7-70cm 29,5 90,8 38,7 52,1 1,177 -0,18 135 26 CV semi-solid<br />

SD17-30cm 20,1 83,8 39,9 43,9 1,451 -0,45 120 24 CV semi-solid<br />

SD17-50cm 21,5 83,9 34,5 49,4 1,263 -0,26 130 25 CV semi-solid<br />

SD25-30cm 21,7 80,1 41,3 38,9 1,503 -0,5 125 25 MV semi-solid<br />

SD25-50cm 21,1 84,4 39,8 44,6 1,419 -0,42 135 25 MV semi-solid<br />

SD29-30cm 17,5 85,3 37,9 47,4 1,429 -0,43 125 27 CV semi-solid<br />

SD29-50cm 21,5 86 35,2 50,8 1,268 -0,27 145 26 CV semi-solid<br />

SD33-30cm 19,8 84,6 37,3 47,3 1,37 -0,37 125 27 CV semi-solid<br />

SD33-50cm 21 85,6 35,4 50,2 1,287 -0,29 145 25 CV semi-solid<br />

SD41-30cm 23,2 84,9 39,3 45,6 1,353 -0,35 130 26 CV semi-solid<br />

SD41-50cm 24,7 90 41,2 48,8 1,338 -0,34 135 26 CV semi-solid<br />

SE1-30cm 23,7 82,2 34,6 47,6 1,229 -0,23 125 23 CV semi-solid<br />

E SE1-50cm 27,1 84,9 37,3 47,6 1,214 -0,21 135 24 CV semi-solid<br />

SE1-70cm 29,9 90,8 39,3 51,5 1,181 -0,18 145 25 CE semi-solid<br />

(black <strong>clays</strong>) SE5-30cm 24,6 84,7 37 47,8 1,258 -0,26 125 24 CV semi-solid<br />

SE5-50cm 26,9 85,5 37,9 47,6 1,23 -0,23 130 25 CV semi-solid<br />

SE5-70cm 29,1 92,2 42,1 50 1,261 -0,26 145 26 MV semi-solid<br />

SE13-30cm 20,8 85,8 39,1 46,7 1,391 -0,39 125 24 CV semi-solid<br />

SE13-50cm 23,5 88,6 39,9 48,7 1,338 -0,34 130 25 CV semi-solid<br />

SE21-30cm 25,2 88,1 39 49,1 1,281 -0,28 125 25 CV semi-solid<br />

SE21-50cm 25,7 90,1 40,7 49,4 1,304 -0,3 145 26 MV semi-solid<br />

SE29-30cm 21,8 80,9 35,5 45,5 1,301 -0,3 120 25 CV semi-solid<br />

SE29-50cm 23 82,5 36 46,5 1,279 -0,28 120 26 CV semi-solid<br />

SE37-30cm 20,8 84,6 37 47,6 1,339 -0,34 125 24 CV semi-solid<br />

SE37-50cm 21,8 85,4 37,8 47,6 1,337 -0,34 130 25 CV semi-solid<br />

SE41-30cm 22 86,7 39,8 46,8 1,381 -0,38 135 25 MV semi-solid<br />

SE41-50cm 20,9 87,5 40 47,5 1,403 -0,4 135 26 MV semi-solid<br />

Rd1 -30cm 39,4 48,6 30,7 17,9 0,512 0,49 20 10 MI s<strong>of</strong>t<br />

Red soils Rd1-100cm 24,2 51,1 30,6 20,5 1,316 -0,32 15 11 MH semi-solid<br />

Rd1-200cm 24,3 49,4 29,6 19,8 1,27 -0,27 20 11 MI-MH semi-solid<br />

Rd1-400cm 25,1 47,7 30 17,7 1,278 -0,28 15 11 MI semi-solid<br />

Rd2-30cm 37,9 53,3 34,7 18,6 0,826 0,17 20 10 MH stiff<br />

Rd2-100cm 26 56,8 34,4 22,4 1,372 -0,37 15 11 MH semi-solid<br />

Rd2-200cm 25,4 54,3 33,2 21,1 1,369 -0,37 15 11 MH semi-solid<br />

Rd2-400cm 26,1 53,3 33,9 19,4 1,401 -0,4 15 11 MH semi-solid


89<br />

Plate 7.2a. Apparatus for liquid limit<br />

determination <strong>of</strong> soil.<br />

Plate7.2b. Liquid limit determination<br />

showing V-shaped groove in soil sample.<br />

The numerical difference beween the liquid limit (LL) <strong>an</strong>d plastic limit (PL) was calculated to<br />

give the plasticity index (PI) <strong>of</strong> the soil, i.e.<br />

PI = LL – PL (7.5)<br />

Results <strong>of</strong> liquid limit tests obtained for the soils in this study are reported to the nearest<br />

whole number (1%) alongside those <strong>of</strong> plastic limit <strong>an</strong>d plasticity index in Table (7.2).<br />

7.1.3.4 Evaluation <strong>an</strong>d application <strong>of</strong> results<br />

Results <strong>of</strong> Atterberg limits obtained in this study have been used for physical classification <strong>of</strong><br />

the clay soils found (Fig. 7.1), based on the st<strong>an</strong>dard plasticity chart or A-line chart (Draft<br />

revision <strong>of</strong> CP 2001; BS 1377: 1975). The clay soils have been classified into one <strong>of</strong> five<br />

categories <strong>of</strong> low plasticity (CL), medium plasticity (CI), high plasticity (CH), very high<br />

plasticity (CV) or extremely high plasticity (CE). Similarly, the silty varieties <strong>of</strong> soils have<br />

been correspondingly classified <strong>an</strong>d denoted by ML, MI, MH, MV, or ME.<br />

In Fig. (7.1), the black <strong>clays</strong> involved in this study have been classified as very high to<br />

extremely high plasticity inorg<strong>an</strong>ic <strong>clays</strong> (CV, CE) <strong>an</strong>d silty <strong>clays</strong> (MV, ME) while the red<br />

soils wholly plot below the A-line <strong>an</strong>d fall in the class <strong>of</strong> medium to high plasticity silts <strong>an</strong>d<br />

silty <strong>clays</strong> (MI, MH).


90<br />

Plasticity Index (PI)<br />

0,5<br />

0,4<br />

0,3<br />

0,2<br />

0,1<br />

0,0<br />

S<strong>of</strong>t<br />

S<strong>an</strong>d<br />

Low<br />

plasticity<br />

(L)<br />

Clays<br />

CL<br />

ML<br />

Mediu<br />

m<br />

plastici<br />

y<br />

High<br />

plasticity<br />

(H)<br />

CH<br />

CI<br />

Silt<br />

MI<br />

MH<br />

BS A-line: PI = 0,73(LL-20)<br />

0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1<br />

Liquid Limit (LL)<br />

Very high<br />

plasticity (V)<br />

CV<br />

MV<br />

Org<strong>an</strong>ic Silt<br />

ME<br />

Black <strong>clays</strong><br />

Red soils<br />

Figure 7.1. A plasticity chart classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study.<br />

On the st<strong>an</strong>dard plasticity chart (BS 1377:1975), the A-line has been defined by the<br />

relationship<br />

PI = 0,73(LL-20) (7.6)<br />

based on experience with British soils (Fig. 7.1).<br />

It was originally drawn from experimental data <strong>an</strong>d serves as a tentative boundary between<br />

clayey <strong>an</strong>d silty varieties <strong>of</strong> cohesive soils which generally plot above <strong>an</strong>d below it,<br />

respectively. For the black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> the present study area, however, the<br />

plasticity index is related to liquid limit through a new relationship derived in this study, i.e.,<br />

PI = 0,79(LL-25) (7.7)<br />

Classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils based on their activity (after Skempton, 1953;<br />

Mitchell, 1976) is presented in Table (7.3). The black <strong>clays</strong> fall in the class <strong>of</strong> normal active<br />

to highly active <strong>clays</strong>, their relatively higher activity being attributed to the higher content <strong>of</strong><br />

smectites (montmorillonites). According to Holtz <strong>an</strong>d Kovacs (1981), montmorillonite has<br />

the highest activity <strong>of</strong> clay minerals <strong>an</strong>d would <strong>of</strong>ten plot above the A-line <strong>of</strong> the plasticity<br />

chart. The red soils are classified as essentially normal active <strong>clays</strong>, <strong>an</strong>d this most probably<br />

due to the predomin<strong>an</strong>t presence <strong>of</strong> kaolinite. Kaolinite is a low activity clay mineral <strong>an</strong>d<br />

would <strong>of</strong>ten plot below the A-line (Holtz <strong>an</strong>d Kovacs, 1981).<br />

The activity state/ level <strong>an</strong>d classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils is also summarised<br />

diagrammatically in Fig. (7.2). From the activity chart, the black <strong>clays</strong> are found to harbour<br />

medium to very high activity levels (i.e. normal to highly active) while the red soils occur in a<br />

medium activity state ( i.e. normal active).


91<br />

Activity chart<br />

Plasticity index PI (%) <strong>of</strong><br />

whole sample<br />

100<br />

50<br />

0<br />

LOW<br />

Activity = 2,0<br />

HIGH<br />

MEDIUM<br />

VERY HIGH<br />

Activity = 1,0<br />

Activity= 0,5<br />

0 50 100<br />

Black <strong>clays</strong><br />

Red soils<br />

% clay in whole sample<br />

Figure 7.2. Activity level classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils in this study (After,<br />

NAVFAC, 1982).<br />

Results <strong>of</strong> Atterberg limits <strong>an</strong>d /or activity values have also been used to asses the probable<br />

type <strong>of</strong> clay minerals present in the soils, based on the classification by Skempton (1953) <strong>an</strong>d<br />

Day (2001), i.e. Table (7.1). It is therefore <strong>an</strong>ticipated that black <strong>clays</strong> which exhibit activity<br />

values <strong>of</strong> 1,10–3,38 (me<strong>an</strong>: 1,87) are mainly smectites (Ca-montmorillonites mainly); while<br />

the red soils with activity values <strong>of</strong> 0,83-1,03 (me<strong>an</strong>: 0,94) would be predominated by<br />

kaolinite <strong>an</strong>d / or illite (Table 7.3). This prediction has been confirmed by actual clay<br />

mineralogical studies on the soils which showed the black <strong>clays</strong> to be over 90% smectites in<br />

composition; while the red soils are mainly kaolinite (80% <strong>an</strong>d over).<br />

In addition, results <strong>of</strong> Atterberg limits obtained during first stage tests served to aid in the<br />

selection <strong>of</strong> samples for chemical/ mineralogical <strong>an</strong>alyses as well as more detailed soil studies<br />

involving shear strength <strong>an</strong>d consolidation-settlement tests.<br />

7.1.4 Free swell test (after Gibbs & Holtz, 1956)<br />

7.1.4.1 Scope<br />

This test was carried out to determine the increase in volume <strong>of</strong> clay soils from a loose dry<br />

powder-form condition when poured into water, expressed as a percentage <strong>of</strong> the original<br />

volume.<br />

According to Gibbs <strong>an</strong>d Holtz (1956), the amount <strong>of</strong> free swell serves to be indicative <strong>of</strong> the<br />

probable swelling <strong>an</strong>d/ or exp<strong>an</strong>sive behaviour <strong>of</strong> clay soils, especially when wetted from<br />

relatively dry conditions under light structural loads.<br />

A classification <strong>of</strong> soils based on free swell is summarised in Table (7.4).


92<br />

Table 7.3. Atterberg limits, activity/ free swell classification <strong>an</strong>d assessment <strong>of</strong> possible clay<br />

mineralogy <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%) PI (%) FS (%) Clay (%) Activity<br />

F. swell<br />

class<br />

Activity class<br />

(BS 1377: 1975)<br />

Possible<br />

clay<br />

minerals<br />

SA1-30cm 32 78 37 41 135 18 2,28 high highly active<br />

A SA1-50cm 30 75 36 39 100 16 2,41 high highly active<br />

SA2-30cm 34 93 39 54 135 26 2,07 high highly active<br />

(black<br />

<strong>clays</strong>) SA2-70cm 34 90 36 54 170 26 2,07 high highly active smectites<br />

SA2-105cm 30 88 33 55 120 24 2,30 high highly active<br />

SA28-30cm 26 85 40 45 125 23 1,96 high active<br />

SA28-50cm 27 89 41 47 135 18 2,62 high highly active<br />

SA37-30cm 19 81 38 43 115 32 1,33 high active<br />

SA37-50cm 22 83 39 44 115 31 1,43 high active<br />

SA37-80cm 23 85 39 46 115 34 1,35 high active<br />

SA41-30cm 23 89 41 47 135 37 1,28 high active<br />

SA41-50cm 25 90 41 49 120 44 1,10 high normal<br />

SB1-30cm 23 82 34 47 125 30 1,58 high active<br />

B SB1-50cm 27 84 36 48 135 26 1,83 high active<br />

SB1-70cm 30 90 39 51 145 25 2,03 high highly active<br />

(black<br />

<strong>clays</strong>) SB5-30cm 24 81 34 47 120 29 1,62 high active smectites<br />

SB5-50cm 27 85 37 48 135 25 1,92 high active<br />

SB5-70cm 31 91 40 51 145 26 1,96 high active<br />

SB7-30cm 23 83 35 48 130 31 1,56 high active<br />

SB7-50cm 26 86 37 49 135 27 1,81 high active<br />

SB7-70cm 30 87 38 49 140 26 1,87 high active<br />

SB18-30cm 20 82 40 43 120 13 3,28 high highly active<br />

SB18-50cm 20 72 34 39 120 14 2,77 high highly active<br />

SB27-30cm 22 85 38 47 130 32 1,47 high active<br />

SB27-50cm 22 88 39 49 130 33 1,47 high active<br />

SB31-30cm 23 87 38 49 130 35 1,39 high active<br />

SB31-50cm 23 86 38 48 130 30 1,61 high active<br />

SB41-30cm 23 89 45 44 145 20 2,21 high highly active<br />

SB41-50cm 24 91 44 47 145 22 2,14 high highly active<br />

SB42-30cm 22 88 44 44 145 23 1,90 high active<br />

SB42-50cm 21 91 45 47 145 21 2,22 high highly active<br />

SC1-30cm 24 82 35 48 125 30 1,58 high active<br />

C SC1-50cm 26 85 37 48 135 27 1,79 high active<br />

SC5-30cm 23 83 35 48 125 31 1,56 high active<br />

(black<br />

<strong>clays</strong>) SC5-50cm 25 84 36 47 135 27 1,75 high active smectites<br />

SC9-30cm 25 84 36 48 130 31 1,55 high active<br />

SC9-50cm 26 85 37 48 135 27 1,76 high active<br />

SC17-30cm 21 76 35 41 105 29 1,41 high active<br />

SC17-50cm 22 81 33 48 110 27 1,79 high active<br />

SC25-30cm 26 88 39 49 125 23 2,13 high highly active<br />

SC25-50cm 26 89 41 48 145 18 2,64 high highly active<br />

SC29-30cm 21 82 32 49 110 28 1,76 high active<br />

SC29-50cm 20 80 37 43 110 26 1,67 high active<br />

SC33-30cm 20 82 33 49 120 27 1,82 high active<br />

SC33-50cm 21 83 33 50 120 28 1,78 high active<br />

SC41-30cm 18 85 38 47 125 16 2,96 high highly active<br />

SC41-50cm 17 84 37 47 125 19 2,47 high highly active


93<br />

Table 7.3 (continued). Atterberg limits, activity/ free swell classification <strong>an</strong>d assessment <strong>of</strong><br />

possible clay mineralogy <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study.<br />

Pr<strong>of</strong>ile<br />

Sample No. Wn (%) LL (%) PL (%) PI (%) FS (%) Clay (%) Activity<br />

F. swell<br />

class<br />

Activity class<br />

(BS 1377:1975)<br />

Possible clay<br />

minerals<br />

SD1-30cm 27 82 36 46 125 32 1,45 high active<br />

D SD1-50cm 27 83 36 47 130 26 1,80 high active<br />

SD1-80cm 30 91 42 49 140 27 1,83 high active<br />

(black <strong>clays</strong>) SD5-30cm 26 84 36 47 125 31 1,52 high active smectites<br />

SD5-50cm 27 84 36 48 130 26 1,84 high active<br />

SD5-80cm 30 92 42 50 140 28 1,80 high active<br />

SD7-30cm 27 84 36 48 110 32 1,50 high active<br />

SD7-50cm 27 86 37 49 140 26 1,88 high active<br />

SD7-70cm 30 91 39 52 135 25 2,08 high<br />

SD17-30cm 20 84 40 44 120 13 3,38 high<br />

highly<br />

active<br />

highly<br />

active<br />

highly<br />

active<br />

SD17-50cm 22 84 35 49 130 21 2,35 high<br />

SD25-30cm 22 80 41 39 125 24 1,62 high active<br />

SD25-50cm 21 84 40 45 135 14 3,19 high<br />

highly<br />

active<br />

SD29-30cm 18 85 38 47 125 27 1,76 high active<br />

SD29-50cm 22 86 35 51 145 28 1,81 high active<br />

SD33-30cm 20 85 37 47 125 29 1,63 high active<br />

SD33-50cm 21 86 35 50 145 28 1,79 high active<br />

SD41-30cm 23 85 39 46 130 35 1,30 high active<br />

SD41-50cm 25 90 41 49 135 41 1,19 high active<br />

SE1-30cm 24 82 35 48 125 31 1,54 high active<br />

E SE1-50cm 27 85 37 48 135 27 1,76 high active<br />

SE1-70cm 30 91 39 52 145 26 1,98 high active<br />

(black <strong>clays</strong>) SE5-30cm 25 85 37 48 125 31 1,54 high active smectites<br />

SE5-50cm 27 86 38 48 130 28 1,70 high active<br />

SE5-70cm 29 92 42 50 145 26 1,92 high active<br />

SE13-30cm 21 86 39 47 125 33 1,42 high active<br />

SE13-50cm 24 89 40 49 130 34 1,43 high active<br />

SE21-30cm 25 88 39 49 125 24 2,05 high<br />

highly<br />

active<br />

SE21-50cm 26 90 41 49 145 20 2,47 high<br />

highly<br />

active<br />

SE29-30cm 22 81 36 46 120 27 1,69 high active<br />

SE29-50cm 23 83 36 47 120 29 1,60 high active<br />

SE37-30cm 21 85 37 48 125 33 1,44 high active<br />

SE37-50cm 22 85 38 48 130 31 1,54 high active<br />

SE41-30cm 22 87 40 47 135 19 2,46 high<br />

highly<br />

active<br />

SE41-50cm 21 88 40 48 135 19 2,50 high<br />

highly<br />

active<br />

Rd1 -30cm 39 49 31 18 20 18 0,99 low normal<br />

Red soils Rd1-100cm 24 51 31 21 15 20 1,03 low normal kaolinite<br />

Rd1-200cm 24 49 30 20 20 24 0,83 low normal<br />

Rd1-400cm 25 48 30 18 15 20 0,89 low normal <strong>an</strong>d/ or<br />

Rd2-30cm 38 53 35 19 20 low<br />

Rd2-100cm 26 57 34 22 15 low illite<br />

Rd2-200cm 25 54 33 21 15 low<br />

Rd2-400cm 26 53 34 19 15 low


94<br />

Table 7.4. Free swell classification <strong>of</strong> clay soils (Gibbs & Holtz, 1956).<br />

FREE SWELL VALUE<br />

FS (%)<br />

FREE SWELL<br />

CLASSIFICATION<br />

POSSIBLE<br />

EXPANSIVENESS<br />

200 very high very high<br />

Up to 2000 (e.g. bentonite) extremely high extremely high<br />

7.1.4.2 Procedure<br />

About 50g <strong>of</strong> a selected soil sample was oven-dried (105 °C) overnight, cooled <strong>an</strong>d passed<br />

through a 400 µm sieve. About 10 ml <strong>of</strong> the sieved dry soil powder was measured out into a<br />

dry 25 ml measuring glass cylinder, without compacting <strong>an</strong>d/ or shaking it down. The soil<br />

powder was then steadily drizzled into 50 ml <strong>of</strong> distilled water contained in a 50 ml glass<br />

measuring cylinder. The main part <strong>of</strong> the solid particles was allowed to come to rest for a<br />

period <strong>of</strong> 30 minutes (Plate 7.3), when the total volume <strong>of</strong> settled solids was read <strong>of</strong>f <strong>an</strong>d<br />

recorded.<br />

Plate 7.3. Free swell determination <strong>of</strong> clay soils.<br />

7.1.4.3 Calculation<br />

Free swell is the ch<strong>an</strong>ge in volume <strong>of</strong> the dry soil expressed as a percentage <strong>of</strong> its original<br />

volume; <strong>an</strong>d is given by<br />

free swell (%) = ((V-Vo)/Vo)*100, or<br />

free swell (%) = ((V-10)/10)*100 (7.8)


95<br />

where<br />

7.1.4.4 Results<br />

V = volume <strong>of</strong> settled solid particles<br />

Vo = 10ml, i.e. original volume <strong>of</strong> dry soil<br />

Results <strong>of</strong> free swell test (FS %) obtained from the present study are given in Table (7.2), <strong>an</strong>d<br />

are reported to the nearest whole number or 1%. Classification <strong>of</strong> soils based on swelling<br />

capabilitiy is summarised in Table (7.3). Black <strong>clays</strong> have high swelling capabilities, being<br />

characterised by free swell values <strong>of</strong> 100% <strong>an</strong>d over. The red soils with free swell values <strong>of</strong><br />

less th<strong>an</strong> 50% (i.e. 15-20%) are classified as low swelling capability <strong>clays</strong>. Potential influence<br />

<strong>an</strong>d/ or effects <strong>of</strong> these swelling capabilities on light constructed structures are discussed in<br />

the next chapter.<br />

7.1.5 Linear shrinkage<br />

7.1.5.1 Scope <strong>an</strong>d procedure<br />

Linear shrinkage tests were carried out in this study according to British St<strong>an</strong>dard testing<br />

procedure (BS 1377: 1975, Test 5). The tests usually serve to determine the percentage linear<br />

shrinkage <strong>of</strong> <strong>clays</strong> as well as soils <strong>of</strong> low plasticity such as silts (Head, 1984). In practice, the<br />

results <strong>of</strong> the tests serve to point to possible volumetric ch<strong>an</strong>ges soils would undergo in situ<br />

during a ch<strong>an</strong>ge <strong>of</strong> seasons from rainy wet months to hot dry ones. In this study, the tests were<br />

performed on both the highly plastic black <strong>clays</strong> as well as the low plasticity silty red soils;<br />

<strong>an</strong>d involved determination <strong>of</strong> linear or one-dimensional ch<strong>an</strong>ge in length <strong>of</strong> a semicylindrical<br />

bar sample <strong>of</strong> clay soil when dried out, starting from near the liquid limit.<br />

The apparatus <strong>an</strong>d procedure are as summarised in Plate (7.4) in which a suitable amount <strong>of</strong><br />

selected soil, previously dried (105 °C) <strong>an</strong>d conveniently prepared, was mixed with distilled<br />

water to produce a homogeneous paste around its liquid limit. The paste, initially filled into a<br />

semi-cylindrical st<strong>an</strong>dard metal (brass) mould <strong>of</strong> 140 mm length <strong>an</strong>d 25 mm diameter, was<br />

dried (105 °C) <strong>an</strong>d the average length (Ld) <strong>of</strong> dried test specimen determined by measuring its<br />

top <strong>an</strong>d lower surface lengths. Linear shrinkage, LS, was calculated as a percentage <strong>of</strong> the<br />

original length <strong>of</strong> the specimen, i.e.<br />

LS (%) = ((Lo-Ld)/Lo)*100 (7.9)<br />

where Lo = original length (140 mm) <strong>of</strong> semi-cylindrical bar sample <strong>of</strong> soil at about its liquid<br />

limit<br />

Ld = average length <strong>of</strong> dried specimen<br />

Results <strong>of</strong> linear shrinkage tests obtained for the soils in this study are presented in Table<br />

(7.2), <strong>an</strong>d are reported to the nearest 1%. Black <strong>clays</strong> exhibit LS values <strong>of</strong> 21-29%, while red<br />

soils show relatively lower values <strong>of</strong> 10-11%.


96<br />

Plate 7.4a. Linear shrinkage determination<br />

showing wet soil in apparatus.<br />

Plate7.4b. Linear shrinkage test:<br />

dried <strong>an</strong>d shrunk soil specimens.<br />

7.1.5.2 Evaluation <strong>of</strong> results<br />

Experience with British soils has shown that the plasticity index, PI, <strong>of</strong> clay soils could be<br />

approximately estimated from results <strong>of</strong> linear shrinkage, LS, through the relationship<br />

PI = 2,13 * LS (7.10)<br />

This relationship is especially useful in soils <strong>of</strong> low clay content, high mica content <strong>an</strong>d low<br />

plasticity; all in which determination <strong>of</strong> Atterberg limits <strong>an</strong>d obtaining <strong>of</strong> reproducible results<br />

are difficult. In such cases, results <strong>of</strong> linear shrinkage tests could be used to give more<br />

consistent plasticity values (BS 1377: 1967).<br />

The application <strong>of</strong> this relationship to results <strong>of</strong> linear shrinkage obtained in this study is<br />

discussed in a later chapter. The relationship between laboratory measured values <strong>of</strong> linear<br />

shrinkage <strong>an</strong>d those <strong>of</strong> plasticity index as obtained for the soils in this study, is also included<br />

for comparison purposes. A correlation <strong>of</strong> laboratory-measured plasticity indices with<br />

calculated plasticity index values (Equation (7.10)) is also given.<br />

Possible <strong>engineering</strong> implications <strong>of</strong> shrinkage behaviour <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils on<br />

construction practice are mentioned in the next chapter.<br />

7.2 Grain/ particle size <strong>an</strong>alysis<br />

7.2.1 Scope<br />

The purpose <strong>of</strong> grain size <strong>an</strong>alysis <strong>of</strong> soils was to group the discrete particles found into<br />

separate r<strong>an</strong>ges <strong>of</strong> sizes, <strong>an</strong>d thereby determine the relative proportions by dry weight <strong>of</strong> each<br />

size r<strong>an</strong>ge. This facilitated to determine whether the soils were predomin<strong>an</strong>tly gravelly, s<strong>an</strong>dy,<br />

silty or clayey; <strong>an</strong>d therefore which <strong>of</strong> these size r<strong>an</strong>ges would most likely control the<br />

<strong>engineering</strong> properties <strong>of</strong> the soils found. The test therefore served to facilitate classification<br />

<strong>of</strong> soils according to particle size, by placing them into one <strong>of</strong> the six arbitrary categories <strong>of</strong><br />

boulders, cobbles, gravel, s<strong>an</strong>d, silt or clay, based on the dominating grain size. As a result,<br />

the <strong>an</strong>alysis served to provide a useful <strong>engineering</strong> classification system whereby soils were<br />

separated into materials <strong>an</strong>d/ or components <strong>of</strong> signific<strong>an</strong>tly differing <strong>engineering</strong> properties.


97<br />

A st<strong>an</strong>dard classification <strong>of</strong> soils by particle size (Glossop <strong>an</strong>d Skempton, 1945; British Code<br />

<strong>of</strong> Practice CP 2001; BS 1377: 1975) is given in Table (7.5), for comparison purposes. In<br />

addition, the clay fraction has been used as <strong>an</strong> index for correlating with other <strong>engineering</strong><br />

properties <strong>of</strong> soils (see Chapter 10).<br />

Application <strong>of</strong> results <strong>of</strong> particle size <strong>an</strong>alysis <strong>of</strong> soils is, however, usually limited by the fact<br />

that <strong>engineering</strong> properties are also influenced by other factors such as mineral type, structure<br />

<strong>an</strong>d <strong>geological</strong> history, <strong>an</strong>d which c<strong>an</strong>not be assessed from particle size tests alone (Head,<br />

1984).<br />

Table 7.5. Soil classification by particle size (BS 1377: 1975).<br />

Particle size (mm) Designation Test procedure<br />

> 200 Boulders Measurement <strong>of</strong><br />

200-60 Cobbles separate pieces<br />

60-20 Coarse gravel<br />

20-6 Medium gravel<br />

6-2 Fine gravel Sieve <strong>an</strong>alysis<br />

2-0,6 Coarse s<strong>an</strong>d<br />

0,6-0,2 Medium s<strong>an</strong>d<br />

0,2-0,06 Fine s<strong>an</strong>d<br />

0,06-0,02 Coarse silt Sedimentation<br />

0,02-0,006 Medium silt Analysis<br />

0,006-0,002 Fine silt _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _<br />

0,002 <strong>an</strong>d less Clay<br />

7.2.2 Procedure <strong>an</strong>d results<br />

Particle size distribution <strong>of</strong> soils in this study was investigated by employing both the sieving<br />

<strong>an</strong>d sedimentation methods <strong>of</strong> <strong>an</strong>alysis. Sieving involved use <strong>of</strong> a series <strong>an</strong>d/ or set <strong>of</strong> 10<br />

sieves <strong>of</strong> st<strong>an</strong>dard aperture openings (Plate 7.5). A wet sieving procedure in accord<strong>an</strong>ce with<br />

British St<strong>an</strong>dard (BS 1377:1975, Test 7(A)) was adopted to facilitate washing down <strong>an</strong>d<br />

separation <strong>of</strong> fines (silt <strong>an</strong>d clay) from the coarse s<strong>an</strong>d <strong>an</strong>d gravel sizes, <strong>an</strong>d thereby determine<br />

the proportions <strong>of</strong> coarse as well as fine material present in the soils<br />

Plate 7.5. Set <strong>of</strong> sieves used for<br />

grain-size <strong>an</strong>alysis <strong>of</strong> soils.<br />

Plate 7.6. Sedimentation apparatus <strong>an</strong>d grain-size<br />

<strong>an</strong>alysis <strong>of</strong> finer fraction <strong>of</strong> soils.


98<br />

The percentage by mass <strong>of</strong> soil retained on each <strong>of</strong> the sieves was calculated, <strong>an</strong>d from this<br />

the cumulative percentage passing (P) each sieve size was then calculated by subtracting the<br />

corresponding summation from 100, i.e.<br />

Pi(%) = 100 - (Σmi/M1) *100 (7.11)<br />

where M1 = total initial dry mass <strong>of</strong> soil<br />

Σmi = cumulative mass retained on upper sieves, down to <strong>an</strong>d including sieve<br />

in question<br />

Pi (%) = cumulative percentage passing sieve size in question<br />

The total mass <strong>of</strong> fines (Mf) <strong>an</strong>d corresponding percentage fines (Pf) representing clay <strong>an</strong>d silt<br />

sizes which pass through the bottom-most (0,063mm) sieve was given by<br />

<strong>an</strong>d<br />

Mf = M1 – Σmi (7.12)<br />

Pf (%)<br />

= ((M1- Σmi)/M1 )*100, or<br />

= 100 – ((Σmi/M1)*100) (7.13)<br />

where M1, Mf <strong>an</strong>d Σmi are as defined above.<br />

The cumulative percentage passing (P%) <strong>of</strong> each sieve size was plotted against corresponding<br />

sieve size (D mm) on a semi-logarithmic chart to give grain size distribution curves <strong>of</strong><br />

<strong>an</strong>alysed soils in the coarse fraction r<strong>an</strong>ge (s<strong>an</strong>d, gravel).<br />

The sedimentation <strong>an</strong>alysis was carried out according to British St<strong>an</strong>dard (BS 1377: 1975,<br />

Test 7(D)), <strong>an</strong>d aimed at investigating the distribution <strong>of</strong> particles in the finer fraction (silt<br />

<strong>an</strong>d clay) <strong>of</strong> soils, by using a suitably calibrated specific gravity hydrometer to monitor <strong>an</strong>d<br />

measure the density <strong>of</strong> a suspension <strong>of</strong> the fraction in water with time (Plate 7.6). The <strong>an</strong>alysis<br />

essentially facilitated the distribution <strong>of</strong> particles in the silt r<strong>an</strong>ge (0,060 – 0,0020 mm) to be<br />

assessed. The maximum diameter <strong>of</strong> particles remaining above a particular depth, H, in the<br />

suspension at <strong>an</strong>y time, t, from the start <strong>of</strong> the sedimentation test was calculated by applying<br />

Stokes´ law (after Sir George Stokes, 1891), i.e.<br />

D = 0,005 531* [ηH/(t(Gs- 1))]^½ (7.14)<br />

where<br />

D = particle diameter (mm)<br />

η = dynamic viscosity <strong>of</strong> water (mPas)<br />

g = acceleration due to gravity (i.e. 9,81m/s²)<br />

H (mm) = effective depth<br />

t (min) = elapsed time<br />

ρw = mass density <strong>of</strong> liquid water (i.e. 1,000 Mg/m³)<br />

Values <strong>of</strong> η <strong>an</strong>d ρw for water temperatures <strong>of</strong> 0-40 °C are given in Table (7.6), i.e. after Kaye<br />

<strong>an</strong>d Laby (1973), while intermediate values may be obtained by either arithmetic or graphical<br />

interpolation.


99<br />

Table 7.6. Viscosity <strong>an</strong>d density <strong>of</strong> water (Kaye <strong>an</strong>d Laby, 1973).<br />

Temperature<br />

(°C)<br />

Dynamic viscosity, η<br />

(mPas)<br />

Density, ρw<br />

(Mg/m³)<br />

0 1,7865 0,999 84<br />

5 1,5138 0,999 95<br />

10 1,3037 0,999 70<br />

15 1,1369 0,999 09<br />

20 1,0019 0,998 20<br />

25 0,8909 0,997 04<br />

30 0,7982 0,995 65<br />

40 0,6540 0,992 22<br />

The relationship in Equation (7.14) is based on the assumption that the particles are<br />

approximately spherical, posses similar densities <strong>an</strong>d exhibit a uniform distribution <strong>an</strong>d small<br />

terminal velocities within the fluid/ water used; <strong>an</strong>d that the fluid/ water used is maintained at<br />

const<strong>an</strong>t temperature, has no turbulence <strong>an</strong>d posseses viscous flow in its still state. It is<br />

therefore implied that large soil particles in a suspension would settle more quickly th<strong>an</strong> small<br />

ones.<br />

The percentage, K, by mass <strong>of</strong> particles smaller th<strong>an</strong> the equivalent diameter, D, was given by<br />

the equation<br />

K (%) = [(Gs * R)/(m(Gs-1))] * 100 (7.15)<br />

where Gs (dimensionless) = specific gravity <strong>of</strong> soil particles<br />

R (dimensionless) = fully corrected hydrometer reading (i.e. ρ-1)*1000<br />

m (g) = mass <strong>of</strong> dry soil; <strong>an</strong>d ρ (g/cm³) = density <strong>of</strong> soil suspension<br />

The values <strong>of</strong> K calculated for each hydrometer reading were plotted against the<br />

corresponding particle size, D (logarithmic scale) to give a particle size distribution curve in<br />

the silt r<strong>an</strong>ge. The intersection <strong>of</strong> this curve with the 0,002 mm ordinate gave the percentage<br />

corresponding to the clay fraction. The curve was plotted on the same semi-logarithmic sheet<br />

as that used for the sieving <strong>an</strong>alysis for the same sample. The two curves were joined to give a<br />

single continouous curve representing the particle size distribution <strong>an</strong>d grading characteristics<br />

<strong>of</strong> the whole sample (Fig. 7.3a). It is observed that grading curves for both black <strong>clays</strong> <strong>an</strong>d red<br />

soils are generally steep <strong>an</strong>d positioned higher up to the left side <strong>of</strong> the chart, i.e. largely<br />

confined within the finer section <strong>of</strong> the chart (Fig. 7.3a), implying that silt <strong>an</strong>d clay are the<br />

major components <strong>of</strong> these soils. The percentages by mass <strong>of</strong> gravel, s<strong>an</strong>d, silt <strong>an</strong>d clay sizes<br />

computed for the soils in this study are presented in Table (7.7).<br />

To aid in classification, the percentage contribution <strong>of</strong> the three major components <strong>of</strong> clay, silt<br />

<strong>an</strong>d coarse fraction (s<strong>an</strong>d + gravel) in each soil sample were plotted in a Tri<strong>an</strong>gular<br />

classification chart (Fig. 7.3b; after BS 1377: 1975). In this chart, the percentage content <strong>of</strong><br />

the three components in <strong>an</strong>y one soil sample would usually add up to 100%. It is evident in<br />

Table (7.7) that s<strong>an</strong>d is generally the predomin<strong>an</strong>t part <strong>of</strong> the coarse fraction in the tested<br />

samples. As result, the terminology “s<strong>an</strong>d” has been adopted in this chart to represent the<br />

combined contribution by the coarse fraction. Classification <strong>of</strong> samples <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red


100<br />

Table 7.7. Results <strong>of</strong> grain size <strong>an</strong>alysis for black <strong>clays</strong> <strong>an</strong>d red soils in this study.<br />

Pr<strong>of</strong>ile<br />

Sample No. Clay (%) Silt (%) S<strong>an</strong>d (%) Gravel (%) % fines<br />

Class<br />

(BS 1377: 1975)<br />

SA1-30cm 18 72 8 2 90 clayey silt<br />

A SA1-50cm 16 56 12 16 72 clayey silt<br />

SA2-30cm 26 64 8 2 90 silty clay with s<strong>an</strong>d<br />

(black <strong>clays</strong>) SA2-70cm 26 62 9 3 88 silty clay with s<strong>an</strong>d<br />

SA2-105cm 24 61 9 6 85 silty clay with s<strong>an</strong>d<br />

SA28-30cm 23 71 5 1 94 silty clay with s<strong>an</strong>d<br />

SA28-50cm 18 70 6 6 88 clayey silt<br />

SA37-30cm 32 57 9 2 89 silty clay<br />

SA37-50cm 31 61 4 4 92 silty clay<br />

SA37-80cm 34 56 8 2 90 silty clay<br />

SA41-30cm 37 59 3 1 96 silty clay<br />

SA41-50cm 44 45 6 5 89 clay<br />

SB1-30cm 30 60 6 4 90 silty clay<br />

B SB1-50cm 26 65 6 3 91 silty clay with s<strong>an</strong>d<br />

SB1-70cm 25 66 5 4 91 silty clay with s<strong>an</strong>d<br />

(black <strong>clays</strong>) SB5-30cm 29 60 7 4 89 silty clay with s<strong>an</strong>d<br />

SB5-50cm 25 66 6 3 91 silty clay with s<strong>an</strong>d<br />

SB5-70cm 26 64 6 4 90 silty clay with s<strong>an</strong>d<br />

SB7-30cm 31 58 7 4 89 silty clay<br />

SB7-50cm 27 64 6 3 91 silty clay with s<strong>an</strong>d<br />

SB7-70cm 26 64 6 4 90 silty clay with s<strong>an</strong>d<br />

SB18-30cm 13 71 8 8 84 clayey silt<br />

SB18-50cm 14 74 9 3 88 clayey silt<br />

SB27-30cm 32 62 4 2 94 silty clay<br />

SB27-50cm 33 59 5 3 92 silty clay<br />

SB31-30cm 35 55 5 5 90 silty clay<br />

SB31-50cm 30 63 5 2 93 silty clay<br />

SB41-30cm 20 68 9 3 88 clayey silt<br />

SB41-50cm 22 64 7 7 86 silty clay with s<strong>an</strong>d<br />

SB42-30cm 23 65 9 3 88 silty clay with s<strong>an</strong>d<br />

SB42-50cm 21 65 8 6 86 silty clay with s<strong>an</strong>d<br />

SC1-30cm 30 57 9 4 87 silty clay<br />

C SC1-50cm 27 63 7 3 90 silty clay with s<strong>an</strong>d<br />

SC5-30cm 31 57 8 4 88 silty clay<br />

(black <strong>clays</strong>) SC5-50cm 27 64 6 3 91 silty clay with s<strong>an</strong>d<br />

SC9-30cm 31 58 7 4 89 silty clay<br />

SC9-50cm 27 62 8 3 89 silty clay with s<strong>an</strong>d<br />

SC17-30cm 29 59 8 4 88 silty clay with s<strong>an</strong>d<br />

SC17-50cm 27 59 6 8 86 silty clay with s<strong>an</strong>d<br />

SC25-30cm 23 71 5 1 94 silty clay with s<strong>an</strong>d<br />

SC25-50cm 18 71 5 6 89 clayey silt<br />

SC29-30cm 28 63 8 1 91 silty clay with s<strong>an</strong>d<br />

SC29-50cm 26 66 7 1 92 silty clay with s<strong>an</strong>d<br />

SC33-30cm 27 64 8 1 91 silty clay with s<strong>an</strong>d<br />

SC33-50cm 28 63 8 1 91 silty clay with s<strong>an</strong>d<br />

SC41-30cm 16 58 8 18 74 clayey silt<br />

SC41-50cm 19 66 8 7 85 clayey silt


101<br />

Table 7.7, continued. Results <strong>of</strong> grain size <strong>an</strong>alysis for black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Pr<strong>of</strong>ile<br />

Sample No. Clay (%) Silt (%) S<strong>an</strong>d (%) Gravel (%) % fines<br />

Class<br />

(BS 1377: 1975)<br />

SD1-30cm 32 59 7 2 91 silty clay<br />

D SD1-50cm 26 62 8 4 88 silty clay with s<strong>an</strong>d<br />

SD1-80cm 27 65 6 2 92 silty clay with s<strong>an</strong>d<br />

(black <strong>clays</strong>) SD5-30cm 31 60 7 2 91 silty clay<br />

SD5-50cm 26 62 8 4 88 silty clay with s<strong>an</strong>d<br />

SD5-80cm 28 64 6 2 92 silty clay with s<strong>an</strong>d<br />

SD7-30cm 32 58 8 2 90 silty clay<br />

SD7-50cm 26 63 7 4 89 silty clay with s<strong>an</strong>d<br />

SD7-70cm 25 66 5 4 91 silty clay with s<strong>an</strong>d<br />

SD17-30cm 13 71 8 8 84 clayey silt<br />

SD17-50cm 21 63 8 8 84 silty clay with s<strong>an</strong>d<br />

SD25-30cm 24 60 8 8 84 silty clay with s<strong>an</strong>d<br />

SD25-50cm 14 60 10 16 74 clayey silt<br />

SD29-30cm 27 63 6 4 90 silty clay with s<strong>an</strong>d<br />

SD29-50cm 28 64 5 3 92 silty clay with s<strong>an</strong>d<br />

SD33-30cm 29 61 6 4 90 silty clay with s<strong>an</strong>d<br />

SD33-50cm 28 63 6 3 91 silty clay with s<strong>an</strong>d<br />

SD41-30cm 35 57 7 1 92 silty clay<br />

SD41-50cm 41 48 6 5 89 clay<br />

SE1-30cm 31 58 7 4 89 silty clay<br />

E SE1-50cm 27 64 6 3 91 silty clay with s<strong>an</strong>d<br />

SE1-70cm 26 65 5 4 91 silty clay with s<strong>an</strong>d<br />

(black <strong>clays</strong>) SE5-30cm 31 58 7 4 89 silty clay<br />

SE5-50cm 28 63 5 4 91 silty clay with s<strong>an</strong>d<br />

SE5-70cm 26 64 6 4 90 silty clay with s<strong>an</strong>d<br />

SE13-30cm 33 60 5 2 93 silty clay<br />

SE13-50cm 34 58 5 3 92 silty clay<br />

SE21-30cm 24 70 5 1 94 silty clay with s<strong>an</strong>d<br />

SE21-50cm 20 68 6 6 88 clayey silt<br />

SE29-30cm 27 65 7 1 92 silty clay with s<strong>an</strong>d<br />

SE29-50cm 29 63 7 1 92 silty clay with s<strong>an</strong>d<br />

SE37-30cm 33 57 5 5 90 silty clay<br />

SE37-50cm 31 61 5 3 92 silty clay<br />

SE41-30cm 19 70 8 3 89 clayey silt<br />

SE41-50cm 19 66 8 7 85 clayey silt<br />

Rd1 -30cm 18 73 9 0 91 clayey silt<br />

Red soils Rd1-100cm 20 74 6 0 94 clayey silt<br />

Rd1-200cm 24 69 7 0 93 silty clay with s<strong>an</strong>d<br />

Rd1-400cm 20 72 8 0 92 silty clay with s<strong>an</strong>d<br />

soils are presented in the chart (Fig. 7.3b). A more summarised classification <strong>of</strong> all tested<br />

samples based on this chart is given in Table (7.7).<br />

In this study therefore, the grading curves provided a me<strong>an</strong>s by which soils could be classified<br />

(Table 7.7) <strong>an</strong>d their <strong>engineering</strong> properties broadly assessed. Engineering implications <strong>of</strong><br />

grain size distribution <strong>an</strong>d grading characteristics <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils on construction<br />

practice are discussed in the next chapter.


102<br />

Particle size distribution curves<br />

100<br />

90<br />

80<br />

70<br />

% passing<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0,001 0,010 0,100 1,000 10,000<br />

Grain size d (mm)<br />

Rd1-100cm (red soils) Rd1-400cm (red soils) SC17-50cm (black <strong>clays</strong>)<br />

SD1-80cm (black <strong>clays</strong>) SE1-30cm (black <strong>clays</strong>) SB18-30cm (black <strong>clays</strong>)<br />

SA2-105cm (black <strong>clays</strong>) SA41-50cm (black <strong>clays</strong>)<br />

Figure 7.3a. Particle size distribution curves for black <strong>clays</strong> <strong>an</strong>d red soils in this study.


103


104<br />

7.3 Direct shear tests<br />

7.3.1 Scope<br />

Direct shear tests are usually carried out to measure <strong>an</strong>d determine the shear strength <strong>of</strong> soils,<br />

by sliding one portion <strong>of</strong> a soil relative to <strong>an</strong>other.<br />

Buildings <strong>an</strong>d other constructed structures impose loads on the soils in <strong>an</strong>d/ or on which they<br />

are founded. The stresses set up in the soil have the effect <strong>of</strong> deforming it through<br />

consolidation by expelling the fluid from the voids between solid particles. The deformation<br />

<strong>of</strong> soil could also occur through shear failure, whereby slippage <strong>of</strong> particles, one on <strong>an</strong>other,<br />

causes the sliding <strong>of</strong> one body <strong>of</strong> soil relative to the surrounding mass. Shear failure is<br />

therefore a result <strong>of</strong> shear stresses set up in the soil exceeding the maximum shear resist<strong>an</strong>ce<br />

the soil c<strong>an</strong> <strong>of</strong>fer, i.e. its shear strength.<br />

Shear strength determinations <strong>of</strong> soils therefore serve to facilitate stability <strong>an</strong>alyses under<br />

various external loading conditions with the purpose <strong>of</strong> guarding against failures which could<br />

cause destruction <strong>of</strong> constructed structures. In practice, economically feasible <strong>an</strong>d safe<br />

constructed structures require that the stresses set up in the soil under loading are everywhere<br />

less th<strong>an</strong> its shear strength by a suitable margin (Colombo, 1983).<br />

According to Head ( 1984), shear strength is usually related to conditions prevailing in situ<br />

<strong>an</strong>d c<strong>an</strong> also vary with time. The value determined in the laboratory is, however, dependent<br />

upon the conditions imposed during the test <strong>an</strong>d, in some inst<strong>an</strong>ces, upon the duration <strong>of</strong> the<br />

test. In this study, the spatial distribution <strong>of</strong> shear strength parameters (c, φ) <strong>of</strong> soils found is<br />

investigated <strong>an</strong>d presented.<br />

7.3.2 Procedure<br />

A shearbox apparatus was used in this study to investigate the shear strength characteristics <strong>of</strong><br />

soils found. The apparatus is relatively quick <strong>an</strong>d simple to use; <strong>an</strong>d consists <strong>of</strong> a rigid metal<br />

box, square in pl<strong>an</strong> with internal dimensions <strong>of</strong> 100 mm length <strong>an</strong>d 30 mm depth, consisting<br />

<strong>of</strong> two halves <strong>an</strong>d in which the soil was placed. The shear test involved sliding the upper<br />

portion <strong>of</strong> the soil contained in upper half <strong>of</strong> the box along <strong>an</strong>d relative to the lower portion in<br />

the lower half, by the action <strong>of</strong> a steadily increasing horizontal shearing displacement (Plate<br />

7.7). A const<strong>an</strong>t load normal to the pl<strong>an</strong>e <strong>of</strong> relative movement was applied at the same time.<br />

During this process, the lower half <strong>of</strong> the box was restrained from moving by a loadmeasuring<br />

device. The horizontal shearing force was applied by a motorised drive unit which<br />

pulled on the upper half <strong>of</strong> the box; while the normal pressure was provided by a yoke<br />

supporting a load h<strong>an</strong>ger.<br />

The area <strong>of</strong> contact between the soil in the two halves <strong>of</strong> the shearbox usually decreases as the<br />

test proceeds, so that a correction to allow for this may be necessary. However, according to<br />

Petley (1966), this affects the shear stress <strong>an</strong>d normal stress in equal proportions so that the<br />

net effect on the failure envelope is usually negligible.


105<br />

Plate 7.7a. Shear testing showing preconsolidation<br />

<strong>an</strong>d shearing <strong>of</strong> specimens.<br />

Plate 7.7b. Shear testing procedure<br />

showing automatic data acquisition.<br />

In this study, slow consolidated-drained (CD) direct shear tests (according to ASTM 3080;<br />

DIN 18 137) were carried out on undisturbed specimens <strong>of</strong> soil samples previously collected<br />

from the field using 100 mm diameter U-tubes. The consolidated-drained conditions were<br />

provided for by allowing soil specimens to completely consolidate under selected normal<br />

pressure before the shearing process to allow for complete dissipation <strong>of</strong> excess pore<br />

pressures. Further drainage was provided for during shear by adjusting the motor speed to<br />

give a suitably slow rate <strong>of</strong> displacement <strong>an</strong>d thereby dissipate <strong>an</strong>y additional pore water<br />

pressure which could have further developed. In practice, the drained shear strength<br />

parameters (Cd, φd) differ only slightly from effective shear strength parameters (c´, φ´ )<br />

obtained from undrained tests in which pore water pressures are measured (Head, 1984). For<br />

m<strong>an</strong>y purposes therefore, the two sets <strong>of</strong> parameters are considered to be equal so that this<br />

study adopts the symbols c´, φ´; for the drained shear strength parameters <strong>of</strong> soils involved.<br />

The shear tests served to provide the <strong>an</strong>gle <strong>of</strong> internal friction, φ (degrees), as well as amount<br />

<strong>of</strong> cohesion, c (kN/m²), <strong>of</strong> the soils investigated. The relative displacement (mm) <strong>of</strong> the two<br />

portions <strong>of</strong> soil specimen as well as the applied shearing force (kN) were measured with time<br />

(t ); <strong>an</strong>d the results used to plot a shear stress/ displacement curve. The vertical movement<br />

(mm) <strong>of</strong> the top surface <strong>of</strong> specimen, which indicates ch<strong>an</strong>ges in volume, was also recorded<br />

<strong>an</strong>d enabled ch<strong>an</strong>ges in density <strong>an</strong>d voids ratio during shear to be evaluated.<br />

Up to three tests at different normal loading conditions were carried out on specimens <strong>of</strong> the<br />

same soil sample <strong>an</strong>d the appropriate shear stress/ displacement curves drawn. The maximum<br />

drained shear resist<strong>an</strong>ce <strong>an</strong>d/ or shear strength <strong>of</strong>fered by the soil at a given normal loading<br />

condition was given by the peak value <strong>of</strong> the respective shear stress/ displacement curve. The<br />

peak or maximum shear stress from each curve was read <strong>of</strong>f <strong>an</strong>d plotted against the<br />

corresponding normal stress to give the failure envelope, i.e. <strong>an</strong> approximately straight linegraph<br />

whose inclination to the horizontal axis represents the <strong>an</strong>gle <strong>of</strong> shearing resist<strong>an</strong>ce, φ.<br />

The cohesion, C, was given by the intercept <strong>of</strong> the line on the vertical stress axis. The straight<br />

line relationship <strong>of</strong> the Coulomb or failure envelope is defined by Coulomb´s law (Coulomb,<br />

1773), i.e.<br />

Tf = C + σnt<strong>an</strong>φ (7.16)<br />

where Tf = maximum shearing resist<strong>an</strong>ce (kN/m²)<br />

σn = normal stress (kN/m²)<br />

C = cohesion (kN/m²)<br />

t<strong>an</strong>φ = frictional component


106<br />

According to Terzaghi <strong>an</strong>d Peck (1967), the cohesion, C, is the component <strong>of</strong> shearing<br />

resist<strong>an</strong>ce due to internal forces holding the particles together in a solid mass; while t<strong>an</strong>φ is<br />

the component <strong>of</strong> shearing resist<strong>an</strong>ce due to interlocking <strong>of</strong> soil particles <strong>an</strong>d friction between<br />

them when subjected to normal stress. The t<strong>an</strong>φ component usually increases with increased<br />

normal stress <strong>an</strong>d disappears under unconfined conditions <strong>of</strong> zero normal loading. On the<br />

other h<strong>an</strong>d, C is independent <strong>of</strong> normal stress <strong>an</strong>d remains const<strong>an</strong>t with increased normal<br />

loading.<br />

7.3.3 Results<br />

The drained shear strength parameters (c´, φ´ ) obtained for the soils in this study are<br />

presented in Table (7.9); with c´ reported to two signific<strong>an</strong>t figures <strong>an</strong>d φ´ to the nearest<br />

degree. The black <strong>clays</strong> exhibit <strong>an</strong>gles <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) <strong>of</strong> between 11° <strong>an</strong>d 30°, giving<br />

a me<strong>an</strong> value <strong>of</strong> 18°. The red soils are characterised by shear resist<strong>an</strong>ce <strong>an</strong>gles <strong>of</strong> 28° to 29°.<br />

These are all typical values <strong>of</strong> clay soils which are usually less th<strong>an</strong> 28° (Lambe <strong>an</strong>d<br />

Whitm<strong>an</strong>, 1979; Johnson <strong>an</strong>d DeGraff, 1988). The characteristic shear stress/ displacement<br />

curves obtained for the soils are included in Fig. (7.4); with the generally cohesive <strong>an</strong>d high<br />

plasticity black <strong>clays</strong> giving rise to sharp peaks while the loose <strong>an</strong>d low plasticity red soils are<br />

characterised by flattened-out peaks. As a result, the difference between peak strength <strong>an</strong>d<br />

residual strength is large in black <strong>clays</strong> <strong>an</strong>d quite insignific<strong>an</strong>t in red soils. A diagrammatic<br />

representation <strong>of</strong> the strength characteristics <strong>of</strong> these soils is illustrated in the form <strong>of</strong> Mohr-<br />

Coulomb failure envelopes in Fig. (7.5).<br />

The very cohesive nature <strong>of</strong> the black <strong>clays</strong> is evidenced by the relatively large values <strong>of</strong><br />

cohesion (c´ ) obtained, i.e.12 – 48 kN/m², giving a me<strong>an</strong> value <strong>of</strong> 35 kN/m² (Table 7.9; Fig.<br />

7.5). On the contrary, cohesion is insignific<strong>an</strong>t in red soils due to their generally loose <strong>an</strong>d<br />

friable nature.<br />

Typical values <strong>of</strong> the <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce for other <strong>geological</strong> materials (quartz grains<br />

<strong>an</strong>d noncohesive soils) are provided in Table (7.8); for comparison purposes.<br />

Table 7.8. Typical values <strong>of</strong> φ for dry noncohesive soils <strong>an</strong>d clay (after Lambe <strong>an</strong>d Whitm<strong>an</strong>,<br />

1979).<br />

Type <strong>of</strong> soil <strong>an</strong>d grading Angle <strong>of</strong> shear resist<strong>an</strong>ce (φ) in degrees<br />

Loose<br />

Dense<br />

Rounded Angular Rounded Angular<br />

Uniform s<strong>an</strong>d- fine to 30 35 37 43<br />

medium<br />

Well-graded s<strong>an</strong>d 34 39 40 45<br />

S<strong>an</strong>d <strong>an</strong>d gravel 36 42 40 48<br />

Gravel 35 40 45 50<br />

Silt 28-32 30-35<br />

Clay < 28 < 30


107<br />

Shear stress / Displacement Curves:<br />

black <strong>clays</strong> <strong>an</strong>d red soils<br />

200,0<br />

180,0<br />

160,0<br />

140,0<br />

Shear stress [kN/m²]<br />

120,0<br />

100,0<br />

80,0<br />

60,0<br />

40,0<br />

20,0<br />

0,0<br />

0,00 5,00 10,00 15,00 20,00 25,00<br />

Horizontal displacement [mm]<br />

SA2-70cm (black <strong>clays</strong>) SB1-30cm (black <strong>clays</strong>) SC1-30cm (black <strong>clays</strong>)<br />

SD7-50cm (black <strong>clays</strong>) RD1-100cm (red soils) RD1-30cm (red soils)<br />

Figure 7.4. Shear stress/ displacement curves <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils in this study.<br />

The variation <strong>of</strong> strength characteristics <strong>of</strong> black <strong>clays</strong> across the study area would be<br />

reflected by the distribution <strong>of</strong> shear strength parameter, (φ`), as shown in Table (7.10). Soil<br />

depths <strong>of</strong> less th<strong>an</strong> 0,50m are characterised by relatively higher maximum (30°), minimum<br />

(16°) <strong>an</strong>d me<strong>an</strong> (21°) values; th<strong>an</strong> those <strong>of</strong> 0,50m depth <strong>an</strong>d greater which have maximum,<br />

minimum <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> 23°, 11° <strong>an</strong>d 17°, respectively. According to Johnson <strong>an</strong>d<br />

DeGraff (1988), the shear strength <strong>of</strong> soils is usually inversely proportional to their plasticity.<br />

As a result, the observed slight decrease <strong>of</strong> strength characteristics <strong>of</strong> black <strong>clays</strong> with depth<br />

could be attributed to a corresponding slight increase <strong>of</strong> their plasticity (PI) with depth (Table<br />

7.10). However, the two depth intervals exhibit more or less similar variation trends <strong>of</strong> soil<br />

strength across the study area as characterised by st<strong>an</strong>dard deviation values which are <strong>of</strong> the<br />

same order <strong>of</strong> magnitude (5,50 <strong>an</strong>d 3,76); <strong>an</strong>d a very strong correlation (R = 0,89) for the two<br />

sets <strong>of</strong> data (Table 7.10). As a result, the observed slight increase in strength parameters (φ`)<br />

with depth could be safely taken as insignific<strong>an</strong>t <strong>an</strong>d the black <strong>clays</strong> therefore classified as<br />

generally homogeneous in their strength characteristics with depth across the study area.


108<br />

Table 7.9. Shear strength parameters <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils. Results <strong>of</strong> natural moisture<br />

content (Wn), plasticity index (PI) <strong>an</strong>d relative consistency (Cr) also included for<br />

comparison purposes.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) PI (%) Cr<br />

Cohesion<br />

c´(kN/m²)<br />

Shear <strong>an</strong>gle<br />

φ´ (degrees)<br />

A SA2-30cm 34 54 1,094 23,10 22<br />

SA2-70cm 34 54 1,04 35,60 19<br />

(black <strong>clays</strong>) SA41-30cm 23 47 1,387 29,80 29<br />

SA41-50cm 25 49 1,334 41,00 19<br />

SB1-30cm 23 47 1,237 46,70 17<br />

B SB1-50cm 27 48 1,197 39,50 12<br />

SB1-70cm 30 51 1,187 17,40 19<br />

(black <strong>clays</strong>) SB5-30cm 24 47 1,212 47,40 17<br />

SB5-50cm 27 48 1,205 36,60 13<br />

SB5-70cm 31 51 1,19 16,00 19<br />

SB7-30cm 23 48 1,235 44,90 17<br />

SB7-50cm 26 49 1,222 36,20 13<br />

SB7-70cm 30 49 1,179 37,60 14<br />

SB41-30cm 23 44 1,491 30,20 30<br />

SB41-50cm 24 47 1,415 28,60 22<br />

SC1-30cm 24 48 1,218 47,20 18<br />

C SC1-50cm 26 48 1,224 31,80 14<br />

SC17-30cm 21 41 1,329 11,50 22<br />

(black <strong>clays</strong>) SC17-50cm 22 48 1,219 33,70 14<br />

SC33-30cm 20 49 1,248 25,10 25<br />

SC33-50cm 21 50 1,233 34,90 22<br />

SD1-30cm 27 46 1,203 47,90 17<br />

D SD1-50cm 27 47 1,19 38,40 16<br />

SD1-80cm 30 49 1,248 33,80 15<br />

(black <strong>clays</strong>) SD5-30cm 26 47 1,217 44,00 16<br />

SD5-50cm 27 48 1,188 36,10 13<br />

SD5-80cm 30 50 1,243 35,10 13<br />

SD7-30cm 27 48 1,198 45,40 16<br />

SD7-50cm 27 49 1,205 36,80 14<br />

SD7-70cm 30 52 1,177 35,50 13<br />

SD33-30cm 20 47 1,37 32,00 30<br />

SD33-50cm 21 50 1,287 28,00 21<br />

SE1-30cm 24 48 1,229 46,40 16<br />

E SE1-50cm 27 48 1,214 39,00 12<br />

SE1-70cm 30 52 1,181 42,50 11<br />

(black <strong>clays</strong>) SE13-30cm 21 47 1,391 42,60 16<br />

SE13-50cm 24 49 1,338 35,10 12<br />

SE29-30cm 22 46 1,301 25,80 26<br />

SE29-50cm 23 47 1,279 36,20 23<br />

SE41-30cm 22 47 1,381 26,00 28<br />

SE41-50cm 21 48 1,403 29,00 22<br />

Rd1 -30cm 39 18 0,512 0,00 29<br />

Red soils Rd1-100cm 24 21 1,316 0,00 29<br />

Rd1-200cm 24 20 1,27 0,00 28<br />

Rd1-400cm 25 18 1,278 0,00 28


109<br />

Table 7.10. Distribution <strong>an</strong>d/ or variation <strong>of</strong> shear strength parameter (φ´) across study area.<br />

Pr<strong>of</strong>ile<br />

Depth (m) Sample No.<br />

Sta. position<br />

(m) PI (%)<br />

Shear <strong>an</strong>gle<br />

(φ´) in<br />

degrees Depth (m) Sample No. PI (%)<br />

SA1-30cm 0 41 SA1-50cm 39<br />

Shear <strong>an</strong>gle<br />

(φ´) in<br />

degrees<br />

A < 0,50 SA2-30cm 250 54 22 ≥ 0,50 SA2-70cm 55 19<br />

SA28-30cm 6750 45 SA28-50cm 47<br />

SA37-30cm 9000 43 SA37-50/80cm 39<br />

SA41-30cm 10000 47 29 SA41-50cm 49 19<br />

SB1-30cm 0 47 17 SB1-50/70cm 50 15<br />

B < 0,50 SB5-30cm 1000 47 17 ≥ 0,50 SB5-50/70cm 50 16<br />

SB7-30cm 1500 48 17 SB7-50/70cm 49 14<br />

SB18-30cm 4250 43 SB18-50cm 39<br />

SB27-30cm 6500 47 SB27-50cm 49<br />

SB31-30cm 7500 49 SB31-50cm 48<br />

SB41-30cm 10000 44 30 SB41-50cm 47 22<br />

SB42-30cm 10250 44 SB42-50cm 47<br />

SC1-30cm 0 48 18 SC1-50cm 48 14<br />

C < 0,50 SC5-30cm 1000 48 ≥ 0,50 SC5-50cm 47<br />

SC9-30cm 2000 48 SC9-50cm 48<br />

SC17-30cm 4000 41 22 SC17-50cm 48 14<br />

SC25-30cm 6000 49 SC25-50cm 48<br />

SC29-30cm 7000 49 SC29-50cm 43<br />

SC33-30cm 8000 49 25 SC33-50cm 50 22<br />

SC41-30cm 10000 47 SC41-50cm 47<br />

SD1-30cm 0 46 17 SD1-50/80cm 48 15<br />

D < 0,50 SD5-30cm 1000 47 16 ≥ 0,50 SD5-50/70cm 49 13<br />

SD7-30cm 1500 48 16 SD7-50/70cm 51 14<br />

SD17-30cm 4000 44 SD17-50cm 49<br />

SD25-30cm 6000 39 SD25-50cm 45<br />

SD29-30cm 7000 47 SD29-50cm 51<br />

SD33-30cm 8000 47 30 SD33-50cm 50 21<br />

SD41-30cm 10000 46 SD41-50cm 49<br />

SE1-30cm 0 48 16 SE1-50/70cm 50 11<br />

E < 0,50 SE5-30cm 1000 48 ≥ 0,50 SE5-50/70cm 49<br />

SE13-30cm 3000 47 16 SE13-50cm 49 12<br />

SE21-30cm 5000 49 SE21-50cm 49<br />

SE29-30cm 7000 46 26 SE29-50cm 47 23<br />

SE37-30cm 9000 48 SE37-50cm 48<br />

SE41-30cm 10000 47 28 SE41-50cm 48 22<br />

max 30 max 22<br />

Statistical < 0,50 min 16 ≥ 0,50 min 11<br />

<strong>an</strong>alysis r<strong>an</strong>ge 14 r<strong>an</strong>ge 11<br />

medi<strong>an</strong> 17 medi<strong>an</strong> 15<br />

mode 17 mode 22<br />

me<strong>an</strong> 21 me<strong>an</strong> 16<br />

st<strong>an</strong>dard dev. 5,50 st<strong>an</strong>dard dev. 3,76<br />

vari<strong>an</strong>ce 30,26 vari<strong>an</strong>ce 14,10<br />

covari<strong>an</strong>ce covari<strong>an</strong>ce -0,08<br />

correlation correlation 0,89


110<br />

Maximum shear stress / Normal stress:<br />

black <strong>clays</strong> & red soils<br />

300,0<br />

250,0<br />

200,0<br />

Shear stress at failure [kN/m²]<br />

150,0<br />

100,0<br />

50,0<br />

0,0<br />

0,0 100,0 200,0 300,0 400,0 500,0 600,0<br />

Normal stress [kN/m²]<br />

SA2-70cm (black <strong>clays</strong>) SB1-30cm (black <strong>clays</strong>) SC1-30cm (black <strong>clays</strong>)<br />

SD7-50cm (black <strong>clays</strong>) RD1-100cm (red soils) RD1-30cm (red soils)<br />

RD1-400cm (red soils)<br />

Figure 7.5. Coulomb/ shear failure envelopes for black <strong>clays</strong> <strong>an</strong>d red soils in this study.<br />

Plotting <strong>of</strong> drained peak values <strong>of</strong> shear stress against normal stress (failure envelopes) was<br />

done on the same diagram, both for black <strong>clays</strong> <strong>an</strong>d red soils (Fig. 7.5).<br />

Possible <strong>engineering</strong> implications <strong>of</strong> shear strength <strong>an</strong>d/ or shear strength parameters <strong>of</strong> soils<br />

on construction practice are discussed under a relev<strong>an</strong>t section in the next chapter.<br />

7.3.4 Evaluation <strong>an</strong>d <strong>an</strong>alysis <strong>of</strong> results<br />

The possible forms <strong>of</strong> relationship between shear parameters (cohesion, <strong>an</strong>gle <strong>of</strong> shear<br />

resist<strong>an</strong>ce) measured on undisturbed samples <strong>an</strong>d index properties <strong>of</strong> black <strong>clays</strong> are


111<br />

illustrated in Figures 7.6, 7.7, 7.8 <strong>an</strong>d 7.9. The index parameters were obtained by performing<br />

tests on disturbed <strong>an</strong>d/ or fractioned samples <strong>of</strong> the <strong>clays</strong>.<br />

Fig. (7.6a) shows the correlation between <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ`) <strong>an</strong>d moisture content<br />

in relation to Atterberg limits, i.e. the relative consistency (Cr), to be fair <strong>an</strong>d <strong>of</strong> moderate<br />

strength (R=0,58). A general increase in the <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce with increased<br />

consistency is observed. However, a broad spread <strong>of</strong> results on the plot diagram occurs, <strong>an</strong>d<br />

serves to be indicative <strong>of</strong> the influence <strong>of</strong> other import<strong>an</strong>t factors such as soil fabric <strong>an</strong>d/ or<br />

structure <strong>an</strong>d clay mineralogy on the magnitude <strong>of</strong> shear <strong>an</strong>gles <strong>an</strong>d, therefore, the strength<br />

characteristics <strong>of</strong> the soils.<br />

Black <strong>clays</strong><br />

Angle <strong>of</strong> shear resist<strong>an</strong>ce<br />

φ´ (°)<br />

35<br />

n = 41<br />

30<br />

25<br />

20<br />

R 2 = 0,34<br />

15<br />

10<br />

1,015 1,215 1,415 1,615<br />

Relative consistency, Cr<br />

Shear <strong>an</strong>gle vs Cr<br />

Figure 7.6a. Correlation between <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) <strong>an</strong>d relative<br />

consistency, Cr, <strong>of</strong> black <strong>clays</strong>.<br />

Black <strong>clays</strong><br />

Cohesion c´(kN/m²)<br />

60,00<br />

n = 41<br />

50,00<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

15 20 25 30 35<br />

Moisture content Wn (%)<br />

Cohesion vs Wn<br />

Figure 7.6b. Correlation between cohesion (c´ ) <strong>an</strong>d moisture content, Wn,<br />

<strong>of</strong> black <strong>clays</strong>.


112<br />

Black <strong>clays</strong><br />

Cohesion c´ (kN/m²)<br />

60,00<br />

50,00<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

n = 41<br />

Cohesion vs Cr<br />

1 1,1 1,2 1,3 1,4 1,5 1,6<br />

Rel. Consistency, Cr<br />

Figure 7.6c. Correlation between cohesion (c´ ) <strong>an</strong>d relative consistency, Cr,<br />

<strong>of</strong> black <strong>clays</strong>.<br />

A poor correlation (R = 0,30) is obtained by relating shear <strong>an</strong>gles <strong>an</strong>d plasticity indices <strong>of</strong> the<br />

<strong>clays</strong> (Fig. 7.7a). A slight improvement in correlation is obtained by plotting the shear <strong>an</strong>gles<br />

against the reciprocal <strong>of</strong> plasticity indices (Fig. 7.8a); but once again, the strength <strong>of</strong><br />

correlation is still low (R = 0,32).<br />

Black <strong>clays</strong><br />

Angle <strong>of</strong> shear resist<strong>an</strong>ce<br />

φ´ (°)<br />

35<br />

n = 41 R² = 0,09<br />

30<br />

25<br />

20<br />

15<br />

10<br />

35,015 40,015 45,015 50,015 55,015 60,015<br />

Plasticity index PI (%)<br />

Shear <strong>an</strong>gle vs PI<br />

Figure 7.7a. Correlation between <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) <strong>an</strong>d plasticity<br />

index (PI) <strong>of</strong> black <strong>clays</strong>.


113<br />

Black <strong>clays</strong><br />

Cohesion c´ (kN/m²)<br />

60,00<br />

n = 41<br />

50,00<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

40 45 50 55<br />

Plasticity index PI (%)<br />

Cohesion vs PI<br />

Fig. 7.7b. Correlation between cohesion (c´ ) & plasticity index (PI) <strong>of</strong> black <strong>clays</strong>.<br />

Black <strong>clays</strong><br />

Angle <strong>of</strong> shear resist<strong>an</strong>ce<br />

φ´ (°)<br />

35<br />

n = 41<br />

30<br />

25<br />

20<br />

R² = 0,102<br />

15<br />

10<br />

0,015 0,020 0,025 0,030<br />

Reciprocal <strong>of</strong> plasticity index 1/P<br />

Shear <strong>an</strong>gle vs 1/PI<br />

Figure 7.8a. Correlation between <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) <strong>an</strong>d reciprocal<br />

<strong>of</strong> plasticity index (1/PI) <strong>of</strong> black <strong>clays</strong>.<br />

Black <strong>clays</strong><br />

Cohesion c´ (kN/m²)<br />

60,00<br />

50,00<br />

n = 41<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

0,018 0,020 0,022 0,024 0,026<br />

1/PI<br />

Cohesion vs 1/PI<br />

Figure 7.8b. Correlation between cohesion (c´ ) <strong>an</strong>d reciprocal <strong>of</strong> plasticity<br />

index (1/PI) <strong>of</strong> black <strong>clays</strong>.


114<br />

The combined influence <strong>of</strong> relative consistency <strong>an</strong>d plasticity indices on the shear <strong>an</strong>gles <strong>an</strong>d<br />

shear strength <strong>of</strong> <strong>clays</strong> is illustrated in Fig. (7.9a). A fair or moderate correlation (R = 0,54)<br />

showing a general increase in shear <strong>an</strong>gles <strong>an</strong>d / or shear strength with <strong>an</strong> increased ratio <strong>of</strong><br />

Cr/PI is observed. There is also <strong>an</strong> improvement in correlation over that based on plasticity<br />

index (Figs. 7.7a & 7.8a) alone.<br />

Black <strong>clays</strong><br />

Angle <strong>of</strong> shear resist<strong>an</strong>ce<br />

φ´ (°)<br />

35<br />

n = 41<br />

30<br />

25<br />

20<br />

R 2 = 0,29<br />

15<br />

10<br />

0,015 0,025 0,035 0,045<br />

Ratio Cr/PI<br />

Shear <strong>an</strong>gle vs Cr/PI<br />

Figure 7.9a. Correlation between <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) <strong>an</strong>d ratio Cr/PI<br />

for black <strong>clays</strong>.<br />

Black <strong>clays</strong><br />

Cohesion c´ (kN/m²)<br />

60,00<br />

50,00<br />

n = 41<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

0,016 0,021 0,026 0,031 0,036<br />

Cr/PI<br />

Cohesion vs Cr/PI<br />

Figure 7.9b. Correlation between cohesion (c´ ) <strong>an</strong>d ratio Cr/PI<br />

for black <strong>clays</strong>.<br />

On the other h<strong>an</strong>d, correlations between values <strong>of</strong> cohesion <strong>an</strong>d those <strong>of</strong> index parameters<br />

(natural moisture content, relative consistency, Atterberg limits) are generally weak, with the<br />

resulting plots giving more spread out data points (Figures 7.6b & c, 7.7b, 7.8b, 7.9b).<br />

All in all, however, the strengths <strong>of</strong> correlations between cohesion <strong>an</strong>d shear <strong>an</strong>gles on one<br />

h<strong>an</strong>d <strong>an</strong>d relative consistency <strong>an</strong>d plasticity indices on the other are generally poor <strong>an</strong>d, at<br />

their best, only moderate. This serves to point to the uncertainities accomp<strong>an</strong>ying assessment<br />

<strong>an</strong>d <strong>characterisation</strong> <strong>of</strong> shear strengths <strong>of</strong> <strong>clays</strong> based on results <strong>of</strong> index tests performed on<br />

disturbed <strong>an</strong>d fractioned / sieved soil samples.


115<br />

7.4 Oedometer consolidation tests<br />

7.4.1 Scope<br />

Oedometer consolidation tests were carried out in this study to investigate <strong>an</strong>d establish<br />

consolidation-settlement characteristics <strong>of</strong> normally consolidated as well as overconsolidated<br />

clay soils found, in terms <strong>of</strong> estimating the amount <strong>an</strong>d rate <strong>of</strong> settlement when externally<br />

loaded. The tests were carried out on saturated <strong>clays</strong>, <strong>an</strong>d were also aimed at determining <strong>an</strong>d<br />

establishing percentage swelling <strong>an</strong>d swelling pressures <strong>of</strong> the clay soils.<br />

Determined parameters <strong>of</strong> coefficient <strong>of</strong> volume compressibility <strong>an</strong>d coefficient <strong>of</strong><br />

consolidation were used to describe the consolidation-settlement characteristics <strong>of</strong> soils. The<br />

coefficient <strong>of</strong> volume compressibility (i.e. modulus <strong>of</strong> volume ch<strong>an</strong>ge) was used to indicate<br />

the compressibility <strong>of</strong> soils in terms <strong>of</strong> the amount by which they would compress when<br />

loaded <strong>an</strong>d allowed to consolidate. The coefficient <strong>of</strong> consolidation is a time related parameter<br />

(Head, 1984), <strong>an</strong>d served to indicate the rate <strong>of</strong> compression <strong>an</strong>d hence the time period over<br />

which consolidation -settlement would take place.<br />

Results <strong>of</strong> the tests <strong>an</strong>d parameters derived there<strong>of</strong> would be usefully applied in foundation<br />

design in limiting settlements to within tolerable limits when structural foundation loads are<br />

imposed on the soils. In addition, the rate <strong>of</strong> settlements <strong>an</strong>d the time within which such<br />

settlements could be virtually complete would be estimated.<br />

The theory <strong>of</strong> consolidation (Terzaghi, 1925; Terzaghi <strong>an</strong>d Peck, 1948; Taylor, 1948) is based<br />

on the fact that, consolidation process <strong>of</strong> soils subjected to external loading mainly entails<br />

escape <strong>of</strong> water from the voids between the skeleton <strong>of</strong> solid grains. The loss <strong>of</strong> pore water is<br />

generally rapid in free-draining materials, but is limited in soils <strong>of</strong> impeded drainage. As a<br />

result, settlements in high permeability materials <strong>of</strong> s<strong>an</strong>ds <strong>an</strong>d gravels generally take place in<br />

a short time, usually as construction work proceeds, thereby causing no major problems. On<br />

the other h<strong>an</strong>d, low permeability <strong>clays</strong> could undergo settlements over much longer periods <strong>of</strong><br />

time after completion <strong>of</strong> construction, thereby subjecting constructed structures to a prolonged<br />

state <strong>of</strong> instability (Nelson <strong>an</strong>d Miller, 1992; Johnson <strong>an</strong>d DeGraff, 1988; Head, 1984).<br />

In this study, the amounts <strong>an</strong>d rates <strong>of</strong> consolidation settlement <strong>of</strong> soils found under various<br />

loading conditions have been studied, <strong>an</strong>alysed <strong>an</strong>d established. The spatial distribution <strong>an</strong>d<br />

variation <strong>of</strong> consolidation parameters across the study area have also been investigated <strong>an</strong>d<br />

established.<br />

7.4.2 Theory <strong>of</strong> consolidation<br />

According to Terzaghi (1926), saturated <strong>clays</strong> which are externally loaded have the total<br />

applied stress, σ, initially supported by the excess pore water pressure, µ, which is induced.<br />

As water drains out <strong>of</strong> the soil, the pore pressure correspondingly falls so that <strong>an</strong> increasing<br />

proportion <strong>of</strong> the applied load is tr<strong>an</strong>sferred to the grains forming the soil skeleton. The<br />

consolidation process therefore involves gradual tr<strong>an</strong>sfer <strong>of</strong> stress from pore water to soil<br />

skeleton. The degree <strong>of</strong> consolidation, U, is a measure <strong>of</strong> the extent <strong>of</strong> this stress tr<strong>an</strong>sfer at<br />

<strong>an</strong>y inst<strong>an</strong>t during consolidation; <strong>an</strong>d is given by the equation<br />

U (%) = ((µo – µ)/ µo) * 100 (7.21)<br />

where µo = initial excess pore pressure


116<br />

µ = excess pore pressure at time t from start <strong>of</strong> consolidation<br />

U = degree <strong>of</strong> consolidation or percentage pore pressure dissipation at time t<br />

The effective stress, σ`, is the numerical difference between total applied stress <strong>an</strong>d pore<br />

water pressure at <strong>an</strong>y inst<strong>an</strong>t during consolidation; <strong>an</strong>d is approximately equal to the stress<br />

carried by the soil skeleton (Simons <strong>an</strong>d Menzies, 1977; Terzaghi, 1926), i.e.<br />

where σ = total applied stress<br />

σ` = effective stress<br />

µ = pore water pressure<br />

σ` = σ – µ (7.22)<br />

However, no measurements <strong>of</strong> pore water pressure are made in the oedometer test so that the<br />

degree <strong>of</strong> consolidation has to be related to the ch<strong>an</strong>ge in height <strong>of</strong> the specimen. By taking<br />

percentage U to be related to the average pore pressure at time t, it c<strong>an</strong> be assumed that the<br />

degree <strong>of</strong> consolidation is proportional to the amount <strong>of</strong> settlement which has taken place by<br />

time t. Thus<br />

U (%) = (∆H/∆Hf) * 100 (7.23)<br />

Where ∆H = settlement up to time t<br />

∆Hf = settlement which will ultimately take place when U = 100%<br />

In practice, the flow <strong>of</strong> water <strong>an</strong>d displacements which take place during consolidation are<br />

nearly always three – dimensional. However, the <strong>an</strong>alysis <strong>of</strong> three-dimensional effects is<br />

extremely complex <strong>an</strong>d rarely practicable (Davis <strong>an</strong>d Poulos, 1965). In this study, Terzaghi´s<br />

one –dimensional <strong>an</strong>alysis was applied in estimating the magnitude <strong>of</strong> settlements <strong>an</strong>d the rate<br />

at which they develop. In applying Terzaghi´s <strong>an</strong>alysis, it would be usually assumed that soils<br />

subjected to consolidation are horizontal, laterally confined, fully saturated, homogeneous <strong>an</strong>d<br />

uniformly thick materials; in which Darcy´s law for the flow <strong>of</strong> water is valid <strong>an</strong>d the<br />

coefficient <strong>of</strong> permeability <strong>an</strong>d other soil properties remain const<strong>an</strong>t during <strong>an</strong>y one increment<br />

<strong>of</strong> applied stress. The <strong>an</strong>alysis also holds for soils in which drainage <strong>an</strong>d compression is onedimensional<br />

in a vertical direction, <strong>an</strong>d soil particles <strong>an</strong>d water are practically incompressible.<br />

In addition, initial excess pore pressure due to application <strong>of</strong> a load should be uniform<br />

throughout the clay layer while one or both <strong>of</strong> adjacent strata to the clay layer are to be<br />

perfectly free-draining in comparison to the clay. The effect <strong>of</strong> the weight <strong>of</strong> the clay layer<br />

under consolidation should be also negligible.<br />

Based on the above assumptions, the simple one-dimensional case <strong>of</strong> consolidation <strong>of</strong> a clay<br />

layer subjected to uniform loading could be summarised in the form <strong>of</strong> a differential equation<br />

(Terzaghi, 1943; Scott, 1974), i.e.<br />

δµ/δt = (K/ρwgmv ) δ²u/dz² (7.24)<br />

where µ = excess pore water pressure at time t, at a given point<br />

z = vertical height <strong>of</strong> point in consideration<br />

K = coefficient <strong>of</strong> permeability <strong>of</strong> clay<br />

mv = coefficient <strong>of</strong> volume compressibility<br />

ρw = mass density <strong>of</strong> water<br />

g = acceleration due to gravity


117<br />

The compound coefficient, K/ρwgmv, is defined as the coefficient <strong>of</strong> consolidation, cv, i.e. a<br />

parameter which relates the ch<strong>an</strong>ge in excess pore pressure with respect to time, to the amount<br />

<strong>of</strong> water draining out <strong>of</strong> the voids <strong>of</strong> a clay layer during the same time, as a result <strong>of</strong><br />

consolidation, or<br />

cv = K/ρwgmv (7.25)<br />

so that Equation (7.24) above becomes<br />

δµ/δt = cv .δ²u/dz² (7.26)<br />

The solution <strong>of</strong> Equation (7.26) yields the percentage consolidation, U, as a function <strong>of</strong> cv, h,<br />

<strong>an</strong>d time t, i.e.<br />

U/100 = f(cvt/h²) (7.27)<br />

where h = length <strong>of</strong> longest drainage path<br />

cvt/h² = a dimensionless number, replaceable by a time factor Tv.<br />

Thus Tv = cvt/h² (7.28)<br />

Equation (7.27) then becomes<br />

U/100 = f(Tv) (7.29)<br />

Computed values <strong>of</strong> U for various values <strong>of</strong> Tv <strong>an</strong>d √Tv are given in Table (7.11) after<br />

Leonards (1962).<br />

Table 7.11. Time factors for one-dimensional<br />

consolidation (Leonards, 1962).<br />

Time factor<br />

U (%)<br />

Tv √Tv<br />

0 0 0<br />

10 0,0077 0,0877<br />

20 0,031 0,176<br />

30 0,071 0,266<br />

40 0,126 0,355<br />

50 0,197 0,443<br />

60 0,286 0,535<br />

70 0,403 0,635<br />

80 0,567 0,753<br />

90 0,848 0,921<br />

95 1,129 1,063<br />

100 ∞ ∞<br />

Fig. 7.10. Time factor Tv related to degree<br />

<strong>of</strong> consolidation U% (Terzaghi, 1943).


118<br />

A graphical relationship between U <strong>an</strong>d Tv is given in Figs. (7.10, 7.11, 7.12) where U is<br />

plotted against Tv, log10(Tv) <strong>an</strong>d √Tv, respectively (Terzaghi, 1943).<br />

Fig. 7.11. Time factor Tv (log scale) related to Fig. 7.12. Square-root <strong>of</strong> time factor, √Tv,<br />

degree <strong>of</strong> consolidation U% related to degree <strong>of</strong> consolidation U%<br />

(Terzaghi, 1943). (Terzaghi, 1943).<br />

The compression <strong>of</strong> soils under a given load is usually divided into three phases (Head, 1988),<br />

<strong>an</strong>d include initial compression, primary consolidation <strong>an</strong>d secondary compression (Fig.<br />

7.13). The initial compression takes place almost simult<strong>an</strong>eously with the application <strong>of</strong> a<br />

load increment in a laboratory test <strong>an</strong>d before commencement <strong>of</strong> drainage. It is a result <strong>of</strong><br />

compression <strong>of</strong> small pockets <strong>of</strong> gas within the pore spaces <strong>an</strong>d partly to bedding down <strong>of</strong><br />

contact surfaces in the cell <strong>an</strong>d load frame. A small proportion may be due to elastic<br />

compression which is recoverable when the load is removed. The initial compression is<br />

usually responsible for deviation <strong>of</strong> the laboratory curve from the theoretical curve near the<br />

beginning <strong>of</strong> a loading increment.<br />

Fig. 7.13. Log-time/settlement curve <strong>an</strong>alysis Fig. 7.14. Analysis <strong>of</strong> square-root time/<br />

method, showing phases <strong>of</strong> consolidation settlement curve (Taylor, 1942).<br />

(Casagr<strong>an</strong>de, 1936).


119<br />

Primary consolidation is a time-dependent compression due to dissipation <strong>of</strong> excess pore<br />

pressure under loading. This phase is accounted for by the Terzaghi consolidation theory<br />

(Terzaghi, 1925), <strong>an</strong>d relates closely to the theoretical curve for most <strong>clays</strong>. It is therefore<br />

used in m<strong>an</strong>y applications to make estimation <strong>of</strong> settlements.<br />

Secondary compression continues after the excess pore pressure <strong>of</strong> the primary consolidation<br />

phase has virtually dissipated, <strong>an</strong>d is most probably a result <strong>of</strong> continued movement <strong>of</strong><br />

particles as the soil structure adjusts itself to increasing effective stress.<br />

The primary consolidation phase is usually used to determine the coefficient <strong>of</strong> consolidation,<br />

cv, for each increment <strong>of</strong> loading through a curve fitting process, which entails comparing a<br />

laboratory consolidation curve with the theoretical curve. In this study, two curve fitting<br />

procedures were adopted <strong>an</strong>d include the log-time method (after Casagr<strong>an</strong>de, 1936), in which<br />

a graphical <strong>an</strong>alysis <strong>of</strong> log-time/ settlement curves was done (Fig. 7.13); <strong>an</strong>d the square-roottime<br />

method (after Taylor, 1942; Barden, 1965) in which a graphical <strong>an</strong>alysis <strong>of</strong> square-roottime/<br />

settlement curves was undertaken (Fig. 7.14).<br />

In this study, the log –time plot was used in the <strong>an</strong>alysis <strong>of</strong> black <strong>clays</strong> whose settlement<br />

curves were generally <strong>of</strong> the conventional shape (Fig. 7.16). However, the square-root plot<br />

was used in the <strong>an</strong>alysis <strong>of</strong> the rather silty red soils, in which otherwise evaluation <strong>of</strong> initial<br />

consolidation conditions, i.e. the do point <strong>of</strong> deformation corresponding to U = 0% primary<br />

consolidation using the log-time method was difficult due to the unconventional shape <strong>of</strong><br />

settlement curves produced (Fig. 7.17).<br />

7.4.3 Consolidation coefficients <strong>an</strong>d other parameters<br />

7.4.3.1 Coefficient <strong>of</strong> consolidation (cv)<br />

This parameter relates the ch<strong>an</strong>ge in excess pore pressure with respect to time, to the amount<br />

<strong>of</strong> water draining out <strong>of</strong> the voids <strong>of</strong> a clay layer during the same time, as a result <strong>of</strong><br />

consolidation.<br />

The coefficient <strong>of</strong> consolidation for a given load increment in a consolidation test c<strong>an</strong> be<br />

determined by re-writing Equation (7.28), i.e.<br />

cv = (Tv/t) * h² (mm²/min) (7.30)<br />

where t (min) = time for a given percentage <strong>of</strong> primary consolidation<br />

h (mm) = length <strong>of</strong> maximum drainage path<br />

A log-time graphical <strong>an</strong>alysis using time t50 corresponding to 50% primary consolidation (i.e.<br />

U = 50%) <strong>an</strong>d theoretical time factor T50 = 0,197 (Table 7.11) tr<strong>an</strong>sforms Equation (7.30) into<br />

cv = (T50/t50) * h²<br />

= 0,197 * (h²/t50)<br />

or cv = (0,197 * (h/1000)²)/(t50/(60*24*365,25)<br />

i.e., cv = 0,1036h²/t50 (m²/year) (7.31)<br />

In the st<strong>an</strong>dard oedometer consolidation test as used in this study, with double drainage to<br />

both the lower <strong>an</strong>d upper surfaces, the height <strong>of</strong> the specimen H' (mm) is equal to 2h.<br />

Therefore Equation (7.31) becomes


120<br />

cv = (0,1036/t50)* (H'/2)²<br />

or cv = 0,026(H')²/t50 (m²/year) (7.32)<br />

where H' (mm) = me<strong>an</strong> specimen height during the load increment<br />

t50 (min) = time for 50% primary consolidation<br />

A square-root time <strong>an</strong>alysis using t90 corresponding to U = 90% primary consolidation <strong>an</strong>d T90<br />

= 0,848 (Table 7.11) tr<strong>an</strong>sforms Equation (7.30) into<br />

cv = (T90/t90)* h²<br />

= 0,848 * (h²/t90) mm²/min, or<br />

cv = (0,848 * (h/1000)²)/(t90/(60*24*365,25))<br />

= 0,446* (h²/t90) m²/year (7.33)<br />

In terms <strong>of</strong> me<strong>an</strong> height <strong>of</strong> specimen, H',<br />

cv = (0,446*(H'/2)²)/t90, or<br />

cv = 0,112 * H'²/t90 (m²/year) (7.34)<br />

where t90 (min) = time for 90% primary consolidation for a given load increment<br />

H' (mm) = me<strong>an</strong> specimen height during this load increment<br />

In practice, however, cv is usually calculated from t50 rather th<strong>an</strong> t90 because the middle<br />

section <strong>of</strong> the laboratory settlement curve is the portion which agrees most closely with the<br />

theoretical curve (Head, 1988).<br />

Typical values <strong>of</strong> coefficient <strong>of</strong> consolidation (Lambe <strong>an</strong>d Whitm<strong>an</strong>, 1979) obtained from<br />

laboratory oedometer tests on specimens <strong>of</strong> uniform soil are related to their approximate<br />

plasticity r<strong>an</strong>ges in Table (7.14). Values <strong>of</strong> cv obtained for black <strong>clays</strong> <strong>an</strong>d red soils in this<br />

study have also been related to their plasticity, <strong>an</strong>d included alongside for comparison<br />

purposes.<br />

7.4.3.2 Voids ratio<br />

Voids ratio, e, is the ratio <strong>of</strong> the volume <strong>of</strong> voids occupied by water <strong>an</strong>d/ or air; to the volume<br />

<strong>of</strong> solid particles in a mass <strong>of</strong> soil.<br />

Voids ratio in a consolidation test is defined by<br />

Voids ratio e = (Gs/ρo) –1 (7.35)<br />

where Gs = specific gravity <strong>of</strong> solid particles<br />

ρ0 = dry density <strong>of</strong> soil (Mg/m³)<br />

Voids ratio, e, is related to degree <strong>of</strong> saturation, S, through the relation<br />

S (%) = (w * Gs/e) * 100 (7.36)


121<br />

where w (%) = moisture content <strong>of</strong> soil<br />

S (%) = degree <strong>of</strong> saturation, i.e. volume <strong>of</strong> water contained in the void space between<br />

soil particles expressed as a percentage <strong>of</strong> total voids<br />

The volume ch<strong>an</strong>ge which occurs during consolidation takes place only in the voids so that a<br />

ch<strong>an</strong>ge in height ∆H from <strong>an</strong> initial height Ho in the specimen corresponds to a ch<strong>an</strong>ge in<br />

voids ratio ∆e from <strong>an</strong> initial value eo. By proportion therefore,<br />

∆H/Ho = ∆e/(1+eo), so that<br />

∆e = ((1+eo)/Ho) * ∆H, or<br />

∆e = F∆H (7.37)<br />

Where F (mm¯¹) = ch<strong>an</strong>ge in voids ratio per unit height, with respect to initial conditions <strong>of</strong><br />

test specimen<br />

The ch<strong>an</strong>ge in height <strong>of</strong> specimen is used to give the voids ratio,e, at <strong>an</strong>y stage <strong>of</strong> a<br />

consolidation test through<br />

e = eo – ∆e (7.38)<br />

where eo = initial voids ratio <strong>of</strong> test specimen<br />

7.4.3.3 Coefficient <strong>of</strong> compressibility (av), coefficient <strong>of</strong> volume compressibility (mv)<br />

<strong>an</strong>d compression index (cc)<br />

These coefficients are derived from consolidation tests <strong>an</strong>d are indicative <strong>of</strong> the<br />

compressibility <strong>of</strong> soils. They serve to give estimates <strong>of</strong> the amount <strong>of</strong> settlement due to<br />

primary consolidation.<br />

The coefficient <strong>of</strong> compressibility, av, <strong>an</strong>d the coefficient <strong>of</strong> volume compressibility, mv, are<br />

usually calculated for each load increment while the compression index, cc, is derived from <strong>an</strong><br />

e/ log p curve obtained by plotting voids ratio, e, against applied pressure, p, (log scale), i.e.<br />

(Fig. 7.15a). In practice, mv is applied to overconsolidated <strong>clays</strong> <strong>an</strong>d cc to normally<br />

consolidated ones. The swell index, cs, gives <strong>an</strong> indication <strong>of</strong> exp<strong>an</strong>sion <strong>of</strong> a soil on<br />

unloading.<br />

The coefficient <strong>of</strong> compressibility, av, is the ch<strong>an</strong>ge in voids ratio, δe, per unit pressure<br />

ch<strong>an</strong>ge, during a given load increment, due to consolidation caused by that incremental<br />

pressure δp, i.e.<br />

av = (e2 – e1)/ δp<br />

= - δe/ δp (m²/kN) (7.39)<br />

where e1 <strong>an</strong>d e2 = voids ratios at start <strong>an</strong>d end <strong>of</strong> consolidation under the load increment<br />

δp in kN/m²<br />

The negative sign indicates decrease <strong>of</strong> e as p increases.


122<br />

Fig. 7.15a. Log-pressure/ voids ratio (e/ log p)<br />

curve .<br />

Fig. 7.15b. Determination <strong>of</strong> coefficient <strong>of</strong><br />

secondary compression cα.<br />

The coefficient <strong>of</strong> volume compressibility, mv, is the ch<strong>an</strong>ge in volume <strong>of</strong> soil test specimen<br />

per unit volume, per unit pressure ch<strong>an</strong>ge, as a result <strong>of</strong> consolidation due to that pressure<br />

ch<strong>an</strong>ge. It is a modulus <strong>of</strong> volume ch<strong>an</strong>ge <strong>an</strong>d serves to indicate the compressibility per unit<br />

thickness <strong>of</strong> soil under test, i.e.<br />

mv = av/(1+e1) = 1/(1+e1) * (- δe/ δp) m²/kN, or<br />

mv = 1000/(1+e1) * (- δe/ δp) m²/MN (7.40)<br />

where e1 = voids ratio at start <strong>of</strong> load increment δp (kN/m²)<br />

Typical values <strong>of</strong> mv for a number <strong>of</strong> types <strong>of</strong> clay are given in Table (7.13), alongside values<br />

<strong>of</strong> mv obtained for black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study, for comparison purposes.<br />

The compression index, cc, is a dimensionless number. It refers to the slope <strong>of</strong> the linear<br />

section <strong>of</strong> a consolidation curve <strong>of</strong> voids ratio, e, plotted against loading pressure, p (log<br />

scale). It is therefore the ch<strong>an</strong>ge in voids ratio for one log cycle <strong>of</strong> pressure ch<strong>an</strong>ge on the<br />

curve whose linear section is represented by the equation<br />

e = eo – cclog10 (po+ δp)/po (7.41)<br />

where eo <strong>an</strong>d po refer to initial conditions <strong>of</strong> voids ratio <strong>an</strong>d pressure.<br />

Skempton (1944), found cc to be related to liquid limit, LL, <strong>of</strong> <strong>clays</strong> to a reasonable degree <strong>of</strong><br />

approximation through<br />

cc = 0,009 (LL-10%) (7.42)<br />

for undisturbed clay, <strong>an</strong>d<br />

cc = 0,007(LL-10%) (7.43)<br />

for remoulded clay.


123<br />

Relationships between values <strong>of</strong> cc <strong>an</strong>d liquid limit for the soils in this study are discussed in a<br />

later section <strong>an</strong>d represented in Table (7.15) <strong>an</strong>d Figs. (7.23 & 7.24). Measured values <strong>of</strong> cc<br />

obtained from consolidation tests in this study are also compared <strong>an</strong>d correlated with<br />

calculated values, both from the new relationship established in this study as well as<br />

Skempton´s relationship (Table 7.16).<br />

7.4.3.4 Swell index (cs)<br />

This is the slope <strong>of</strong> the swelling or unloading curve <strong>of</strong> e plotted against log p (Fig. 7.18). It is<br />

the ch<strong>an</strong>ge in voids ratio, e, over one log cycle <strong>of</strong> pressure ch<strong>an</strong>ge in the linear section <strong>of</strong> the<br />

swelling/ unloading curve <strong>of</strong> e/ log p. The straight-line section <strong>of</strong> the curve is given<br />

approximately by the equation<br />

e = eo ± cslog10 (po+δp)/po (7.44)<br />

where eo <strong>an</strong>d po refer to initial conditions <strong>of</strong> voids ratio <strong>an</strong>d loading pressure.<br />

Correlations between swell index <strong>an</strong>d index properties are discussed in a later chapter.<br />

7.4.3.5 Coefficient <strong>of</strong> permeability (K)<br />

The coefficient <strong>of</strong> permeability, K, is usually computed from Equation (7.25), i.e.<br />

cv = K/(ρwgmv ), or by re-writing<br />

K = cvmvρwg (mm/s), or in practical units<br />

K = cv/(365,25*24*3600) * (mv*10^¯6) * (1*10³) * 9,81 (m/s), or<br />

K = cvmv * 0,31*10^¯9 (m/s) (7.45)<br />

where cv (m²/year) <strong>an</strong>d mv (m²/MN) are consolidation parameters whose values are known,<br />

g (m/s²) = acceleration due to gravity<br />

ρw (kg/m³) = mass density <strong>of</strong> water<br />

Values <strong>of</strong> K calculated for black <strong>clays</strong> <strong>an</strong>d red soils <strong>of</strong> this study using the above equation are<br />

given in Table (7.12).<br />

7.4.3.6 Coefficient <strong>of</strong> secondary compression (cα)<br />

This is the compression which continues after primary consolidation has virtually finished,<br />

<strong>an</strong>d is time dependent. According to Head (Head, 1988), it is a signific<strong>an</strong>t factor in the<br />

settlement <strong>of</strong> s<strong>of</strong>t soils, especially org<strong>an</strong>ic <strong>clays</strong> <strong>an</strong>d peats where it usually increases with<br />

increased load application.<br />

Secondary compression is equal to the slope <strong>of</strong> the linear portion <strong>of</strong> the secondary<br />

compression part <strong>of</strong> log-time/ settlement plot, in terms <strong>of</strong> strain per log cycle <strong>of</strong> time (Fig.<br />

7.15b), i.e.<br />

cα = (δHs/Ho)/ δlogt, or for one log cycle <strong>of</strong> time ch<strong>an</strong>ge


124<br />

cα = δHs/Ho (7.46)<br />

where δHs (mm) = ch<strong>an</strong>ge in specimen height over one log-cycle <strong>of</strong> time<br />

Ho (mm) = initial height <strong>of</strong> specimen<br />

cα = a dimensionless number<br />

R<strong>an</strong>ges <strong>of</strong> values <strong>of</strong> cα determined for black <strong>clays</strong> <strong>an</strong>d red soils in this study are presented <strong>an</strong>d<br />

compared with typical values for other <strong>clays</strong> in Table (7.17).<br />

7.4.4 Procedure<br />

Oedometer consolidation tests were carried out in this study according to British St<strong>an</strong>dard (BS<br />

1377: 1975, Test (17)). The tests were carried out on undisturbed specimens <strong>of</strong> the soils<br />

previously collected from the project area. Use was made <strong>of</strong> a fixed-ring type <strong>of</strong> consolidation<br />

cell consisting <strong>of</strong> a mould <strong>an</strong>d/ or steel cutting ring (for rigidly supporting the soil specimen),<br />

upper <strong>an</strong>d lower drainage surfaces, a loading cap <strong>an</strong>d a provision for adding <strong>an</strong>d/ or<br />

containing distilled water to keep <strong>an</strong>d maintain the soil specimen saturated. The ring was<br />

capable <strong>of</strong> accommodating specimens <strong>of</strong> 14 mm thickness <strong>an</strong>d 71,40 mm diameter. According<br />

to British St<strong>an</strong>dard (BS 1377: 1975), the height to diameter ratio <strong>of</strong> specimens is usually<br />

about 1: 4 to 1:5. The small thickness <strong>of</strong> test specimens used served to ensure that testing<br />

times were not excessively long; <strong>an</strong>d the tests could also be easily extended to long-term tests<br />

to provide secondary compression characteristics <strong>of</strong> soils. The set-up <strong>of</strong> apparatus as used in<br />

this study is presented in Plate (7.8).<br />

A st<strong>an</strong>dard sequence <strong>of</strong> vertical loads r<strong>an</strong>ging from 6 to 2400 kPa was applied to the laterally<br />

confined specimen contained in the consolidation ring; with 6-12 kPa as extended r<strong>an</strong>ge for<br />

very s<strong>of</strong>t soils; 25-800 kPa as normal r<strong>an</strong>ge; <strong>an</strong>d 1600-2400 kPa as extended r<strong>an</strong>ge for stiff to<br />

hard <strong>an</strong>d/ or overconsolidated soils. The actual initial incremental load used for the black<br />

<strong>clays</strong> depended on the amount <strong>of</strong> swelling pressure exhibited by the soils. The resulting<br />

vertical compression (mm) under each load was observed <strong>an</strong>d recorded over a 24 hour period<br />

<strong>of</strong> time. The laterally confining ring had the purpose <strong>of</strong> permitting vertical deformation <strong>of</strong> the<br />

soil specimen, without lateral deformation, so that one-dimensional consolidation<br />

characteristics <strong>of</strong> soils could be investigated.<br />

At the end <strong>of</strong> the last incremental loading stage, the specimen was unloaded in a sequence <strong>of</strong><br />

decrements <strong>an</strong>d allowed to swell in about half the number <strong>of</strong> stages as were applied during<br />

consolidation. Compression readings were plotted cumulatively against time, during both<br />

loading (consolidation) <strong>an</strong>d unloading (swelling) stages (Figs. 7.16 & 7.17). Compression/<br />

time curves obtained for loading stages were used for derivation <strong>of</strong> one-dimensional<br />

consolidation parameters.<br />

Subsequent calculations involving consolidation test data made use <strong>of</strong> <strong>an</strong> average value <strong>of</strong><br />

specific gravity <strong>of</strong> soil particles <strong>of</strong> 2,72 for red soils <strong>an</strong>d 2,70 for black <strong>clays</strong>; assuming<br />

inorg<strong>an</strong>ic <strong>clays</strong> for the two types <strong>of</strong> soil.


125<br />

Plate 7.8. Oedometer consolidation testing in this study.<br />

The British St<strong>an</strong>dard consolidation testing procedure adopted in this study is very similar in<br />

principle to the one-dimensional consolidation test specified by the ASTM Designation<br />

D2435. However, the ASTM st<strong>an</strong>dard incorporates either a fixed-ring type or floating-ring<br />

type <strong>of</strong> consolidation cell, with minimum specimen dimensions <strong>of</strong> 50 mm diameter <strong>an</strong>d 12,5<br />

mm thickness; <strong>an</strong>d a minimum height to diameter ratio <strong>of</strong> 1:2,5. In addition, a copper disc is<br />

also used for the deformation calibration <strong>of</strong> the apparatus, while <strong>an</strong> initial seating load <strong>of</strong> 2,5-<br />

5,0 kPa is usually applied to the soil specimen depending on the s<strong>of</strong>tness <strong>of</strong> the soil. In<br />

practice, a st<strong>an</strong>dard loading sequence <strong>of</strong> pressures consist <strong>of</strong> 5, 12,5; 25, 50, 100, 200, 400,<br />

800 kPa, even though smaller increments could be used on very s<strong>of</strong>t soils (BS 1377: 1975;<br />

ASTM D2435-70).<br />

7.4.5 Analysis <strong>of</strong> consolidation test data<br />

A graphical <strong>an</strong>alysis <strong>of</strong> consolidation test data obtained for black <strong>clays</strong> made use <strong>of</strong><br />

conventional log-time method. However, both log-time <strong>an</strong>d square-root-time methods were<br />

employed in the graphical <strong>an</strong>alysis <strong>of</strong> test data for the more silty red clay soils, whose early<br />

portion <strong>of</strong> log- time/ settlement curves depart from the conventional shape as derived from<br />

theoretical relationships (Fig. 7.17). Determination <strong>of</strong> the do point corresponding to 0%<br />

primary consolidation was therefore not possible. The departure from conventional shape<br />

could be attributed to the relatively much higher permeability (k = 7,57E-11 to 7,33E-9) <strong>of</strong><br />

the red soils which caused rapid settlement immediately after initial loading, so that the initial


126<br />

convex-upwards portion <strong>of</strong> the primary curve was passed before <strong>an</strong>y readings could be taken.<br />

In this case, the do point was estimated from the square-root-time/ settlement curve <strong>an</strong>d then<br />

tr<strong>an</strong>sferred to the log-time curve for the conventional <strong>an</strong>alysis to be carried out. According to<br />

Head (1988), cv values in the r<strong>an</strong>ge <strong>of</strong> 10-100 m²/year are indicative <strong>of</strong> quite rapid settlements<br />

which would not be expected to cause long-term instability problems.<br />

Some <strong>of</strong> the red soils classified as clayey silts to silt types gave rise to log-time/ settlement<br />

curves which were concave upwards from the start implying relatively rapid draining soils.<br />

Both the initial convex portion <strong>an</strong>d point <strong>of</strong> inflexion (Figs. 7.11 & 7.13) were passed before<br />

the first recorded reading was made so that determination <strong>of</strong> do <strong>an</strong>d d100 points corresponding<br />

to 0% <strong>an</strong>d 100% primary consolidation, respectively, was difficult. The square-root-time<br />

curve could also not be used for determination <strong>of</strong> the do point due to lack <strong>of</strong> the necessary<br />

linear portion <strong>of</strong> the curve (Figs. 7.12 & 7.14). The value <strong>of</strong> cv was therefore estimated by<br />

assuming the occurrence <strong>of</strong> the d50 point somewhere mid-way between points do <strong>an</strong>d df on the<br />

log-time curve (Fig. 7.13). A value <strong>of</strong> t50 was then estimated <strong>an</strong>d used to calculate cv. A r<strong>an</strong>ge<br />

<strong>of</strong> t50 less th<strong>an</strong> 0,10 minutes, implied<br />

cv > (0,0256 * H'²)/t50 = 0,0256 * 14²/0,1 or<br />

cv > 50 m²/year, implying very rapid drainage <strong>an</strong>d settlement <strong>of</strong> the soil,<br />

where H' (mm) = me<strong>an</strong> specimen height<br />

d50 (mm) = compression reading for 50% primary consolidation<br />

dc (mm) = corrected initial compression reading<br />

df (mm) = final compression reading<br />

t5o (min) = time for 50% primary consolidation<br />

Use <strong>of</strong> a Rowe consolidation cell (Rowe,1966), capable <strong>of</strong> accommodating larger soil<br />

specimens would be recommended for oedometer tests on the clayey silt <strong>an</strong>d silty varieties <strong>of</strong><br />

red soils, so as to obtain a more definite value <strong>of</strong> cv for estimation <strong>of</strong> rates <strong>of</strong> settlement under<br />

various loads.<br />

7.4.6 Results <strong>of</strong> consolidation tests<br />

Results <strong>of</strong> log-time settlement curves obtained for black <strong>clays</strong> <strong>an</strong>d red soils in this study are<br />

illustrated in Figures (7.16) <strong>an</strong>d (7.17), respectively. The red soil specimen is shown to<br />

undergo settlements <strong>of</strong> more th<strong>an</strong> 6mm over a loading r<strong>an</strong>ge <strong>of</strong> up to 2400 kPa. Settlements<br />

undergone by black clay specimen are less th<strong>an</strong> 3mm over the same loading r<strong>an</strong>ge. Results in<br />

this study therefore show the red soils to be relatively more compressible th<strong>an</strong> the black <strong>clays</strong>.<br />

In addition, unloading curves for the black <strong>clays</strong> are more inclined th<strong>an</strong> those <strong>of</strong> the red soils<br />

which are more or less flat. This is indicative <strong>of</strong> the black <strong>clays</strong> having relatively more swelling<br />

capabilities <strong>an</strong>d/ or exp<strong>an</strong>sion tendencies on unloading th<strong>an</strong> the red soils.


127<br />

Cumulative log-time/settlement<br />

curves: black <strong>clays</strong> (SA2-70cm)<br />

Cumulative log-time/settlement<br />

curves: Red soils (RD1-100cm)<br />

cumulative settlement (mm)<br />

time - minutes (log scale)<br />

0,1 10,0 1000,0 100000,0<br />

0<br />

0,2<br />

0,4<br />

0,6<br />

0,8<br />

1<br />

1,2<br />

1,4<br />

1,6<br />

1,8<br />

2<br />

2,2<br />

cumulative settlement (mm)<br />

0,100 10,000 1000,000<br />

0,000<br />

1,000<br />

2,000<br />

3,000<br />

4,000<br />

5,000<br />

time - minutes (log scale)<br />

100000,0<br />

00<br />

2,4<br />

2,6<br />

6,000<br />

2,8<br />

3<br />

Fig. 7.16. Conventional log-time/settlement<br />

curves for black <strong>clays</strong> during loading<br />

<strong>an</strong>d unloading stages.<br />

7,000<br />

Fig. 7.17.Unconventional log-time/settlement<br />

curves for red soils during loading<br />

<strong>an</strong>d unloading stages.<br />

Results <strong>of</strong> voids ratio ch<strong>an</strong>ge <strong>of</strong> soils during loading <strong>an</strong>d unloading stages are summarised in<br />

the form <strong>of</strong> voids ratio/ log pressure (e/log p) curves as shown in Figure (7.18). The curves<br />

for the red soils are generally placed higher up the diagram, implying higher voids ratios <strong>an</strong>d<br />

porosities. On the other h<strong>an</strong>d, e/log p curves for black <strong>clays</strong> occupy lower positions on the<br />

diagram due to their relatively lower voids ratios <strong>an</strong>d porosities. As a result, red soils would<br />

be expected to be more compressible th<strong>an</strong> the black <strong>clays</strong> over the same loading r<strong>an</strong>ge. In<br />

addition, unloading portions <strong>of</strong> the curves are more or less flat <strong>an</strong>d/ or horizontal for the red<br />

soils, while those for black <strong>clays</strong> are relatively steeper. This serves to show once again the<br />

higher potential swelling capabilities <strong>of</strong> black <strong>clays</strong> relative to those <strong>of</strong> red soils.


128<br />

Consolidation curves <strong>of</strong> e/log p<br />

Voids ratio (e)<br />

1,501<br />

1,401<br />

1,301<br />

1,201<br />

1,101<br />

1,001<br />

0,901<br />

0,801<br />

0,701<br />

0,601<br />

0,501<br />

0,401<br />

0,301<br />

0,201<br />

0,101<br />

0,001<br />

10 100 1000 10000<br />

Pressure on log scale (log P (kPa))<br />

RD1-30cm: red soils RD1-100cm: red soils RD1-400cm: red soils<br />

SA2-70cm: black <strong>clays</strong> SA37-50cm: black <strong>clays</strong> SB1-70cm: black <strong>clays</strong><br />

SB42-50cm: black <strong>clays</strong> SC17-50cm: black <strong>clays</strong> SC29-50cm: black <strong>clays</strong><br />

Figure 7.18. Voids ratio/ log Pressure (e/ log p) curves for black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Results <strong>of</strong> oedometer consolidation tests in terms <strong>of</strong> voids ratio, compression coefficients <strong>an</strong>d/<br />

or indices as well as permeability over the loading r<strong>an</strong>ge <strong>of</strong> 25-800 kPa are summarised in<br />

Table (7.12). Complete results for the loading <strong>an</strong>d unloading stages <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils<br />

are tabulated in Appendix A.<br />

Table 7.12. Derived parameters from results <strong>of</strong> oedometer consolidation tests for the loading<br />

r<strong>an</strong>ge <strong>of</strong> 25-800 kPa in this study.<br />

Sample<br />

t50 Cv<br />

No. E (%) e av (m²/kN) (min) (m²/year) K (m/s) Cc Cs Cα<br />

mv<br />

(m²/MN)<br />

SA2-70cm 0,31-11,11 0,52-0,70 0,000052-0,00045 0,03-0,26 1,00 - 60 0,07-5,08 2,6E-12 - 4,8E-11 0,3 0,07 0,0002-0,009<br />

SA37-50cm 1,84-27,68 0,52-1,06 0,0003-0,0033 0,18-1,62 2,90-55 0,05-1,73 2,94E-12 - 3,94E-10 0,4 0,075 0,003-0,017<br />

SB1-70cm 1,85-19,46 0,65-1,01 0,000148-0,0015 0,09-0,74 0,48-18 0,19-2,36 5,11E-12 - 1,2E-9 0,29 0,043 0,0002-0,006<br />

SB42-50cm 8,02-29,11 0,23-0,59 0,0003-0,0014 0,22-0,80 9,5-50 0,06-0,494 3,92E-12 - 1,23E-10 0,34 0,03 0,005-0,024<br />

SC17-50cm 1,74-20,64 0,37-0,7 0,00028-0,0010 0,17-0,60 9,0- 50 0,07-0,56 1,67E-12 - 3,0E-11 0,44 0,04 0,006-0,016<br />

SC29-50cm 3,33-18,24 0,56-0,84 0,00026-0,0008 0,16-0,44 2,5 -35 0,104-1,97 5,03E-12 - 2,03E-10 0,38 0,065 0,003-0,13<br />

RD1-30cm 10,98-32,07 0,74-1,28 0,00035-0,0047 0,19-2,0 0,35-1,85 1,38-11,85 7,99E-11 - 7,33E-9 0,42 0,03 0,003-0,004<br />

RD1-100cm 7,35-39,01 0,59-1,42 0,00036-0,0074 0,21-2,94 0,98-1,75 1,18-4,67 7,57E-11 - 4,26E-9 0,4 0,03 0,005-0,008<br />

RD1-400cm 3,09-19,60 0,74-1,1 0,00029-0,0024 0,16-1,11 0,40-1,30 2,71-12,14 1,32E-10 - 4,19E-9 0,39 0,02 0,002-0,007


129<br />

R<strong>an</strong>ges <strong>of</strong> values <strong>of</strong> coefficient <strong>of</strong> volume compressibility, mv, obtained for black <strong>clays</strong> <strong>an</strong>d<br />

red soils are given in Table (7.13), where they have been used to classify the soils based on<br />

compressibility. R<strong>an</strong>ges <strong>of</strong> values <strong>of</strong> mv for other types <strong>of</strong> soil are also included for<br />

comparison purposes. Results in Table (7.13) show that black <strong>clays</strong> would generally exhibit<br />

medium to high compressibility over the normal loading r<strong>an</strong>ge <strong>of</strong> 25-800kPa; however, cases<br />

<strong>of</strong> low to medium compressibility are also to be expected. The red soils would be<br />

characterised by medium to very high compressibility over the same loading r<strong>an</strong>ge.<br />

Comparatively, therefore, the red soils tend to be more compressible th<strong>an</strong> the black <strong>clays</strong>.<br />

This difference could be attributed to the generally loose <strong>an</strong>d friable nature <strong>of</strong> the red soils<br />

giving rise to higher porosities, as evidenced by voids ratios (e) <strong>of</strong> 0,59-1,42 (Table 7.12).<br />

Lower compressibility <strong>of</strong> the black <strong>clays</strong> could be attributed to lower porosities (e = 0,23-<br />

1,06), a most probable result <strong>of</strong> the cohesive nature <strong>an</strong>d dense compact structure <strong>of</strong> these<br />

<strong>clays</strong>.<br />

Table 7.13. Compressibility classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils in this study.<br />

Coeff. <strong>of</strong> volume<br />

Category<br />

Clay type/ Sample No.<br />

compressibility mv<br />

(m²/MN)<br />

Compressibility<br />

classification<br />

Typical Very org<strong>an</strong>ic alluvial <strong>clays</strong> > 1,5 very high<br />

examples (after <strong>an</strong>d peats<br />

Lambe & Whitm<strong>an</strong> Normally consolidated 0,3-1,5 high<br />

-1979) alluvial/ estuarine <strong>clays</strong><br />

Fluvio-glacial <strong>clays</strong> <strong>an</strong>d 0,1-0,3 medium<br />

lake <strong>clays</strong><br />

Boulder <strong>clays</strong> 0,05-0,1 low<br />

Heavily overconsolidated < 0,05 very low<br />

boulder <strong>clays</strong><br />

SA2-70cm 0,03-0,26 very low to medium<br />

This study: SA37-50cm 0,18-1,62 medium to very high<br />

Black <strong>clays</strong> SB1-70cm 0,09-0,74 low to high<br />

SB42-50cm 0,22-0,80 medium to high<br />

SC17-50cm 0,17-0,60 medium to high<br />

SC29-50cm 0,16-0,44 medium to high<br />

RD1-30cm 0,19-2,0 medium to very high<br />

This study: RD1-100cm 0,21-2,94 medium to very high<br />

Red soils RD1-400cm 0,16-1,11 medium to high<br />

R<strong>an</strong>ges <strong>of</strong> values <strong>of</strong> coefficient <strong>of</strong> consolidation, cv, obtained for the black <strong>clays</strong> <strong>an</strong>d red soils<br />

are given in Table (7.14). Typical r<strong>an</strong>ges for other inorg<strong>an</strong>ic soils are included for comparison<br />

purposes. From the results, the red soils would undergo medium to high rates <strong>of</strong> consolidation<br />

–settlement when externally loaded in the r<strong>an</strong>ge <strong>of</strong> 25-800 kPa. The rates <strong>of</strong> consolidationsettlement<br />

for the black <strong>clays</strong> would be generally low over the same loading r<strong>an</strong>ge. The high<br />

rates <strong>of</strong> consolidation by the red soils could be attributed to their relatively high porosity <strong>an</strong>d<br />

high permeability (K = 7,57*E-11 to 7,33*E-9 m/s ) which facilitate faster drainage <strong>an</strong>d rapid<br />

dissipation <strong>of</strong> pore water pressures. On the other h<strong>an</strong>d, the black <strong>clays</strong> are characterised by<br />

relatively lower porosity <strong>an</strong>d permeability (K = 1,67*E-12 to 1,2*E-9 m/s) so that drainage<br />

<strong>an</strong>d dissipation <strong>of</strong> pore pressures on external loading would be less rapid, causing slower<br />

consolidation –settlements.


130<br />

Table 7.14. Rates <strong>of</strong> consolidation-settlement <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Category<br />

Clay type/ Sample No.<br />

Coeff. <strong>of</strong> consolidation cv<br />

(m²/year)<br />

Rate <strong>of</strong> consolidationsettlement<br />

Typical examples High plasticity 0,1-1,0 low<br />

for inorg<strong>an</strong>ic <strong>clays</strong> Medium plasticity 1,0- 10 Medium<br />

(after Lambe & Low plasticity 10-100 High<br />

Whitm<strong>an</strong> –1979) Silts > 100 very high<br />

SA2-70cm 0,07-5,08 low - medium<br />

This study: SA37-50cm 0,05-1,73 low -medium<br />

Black <strong>clays</strong> SB1-70cm 0,19-2,36 low -medium<br />

SB42-50cm 0,06-0,494 low<br />

SC17-50cm 0,07-0,56 low<br />

SC29-50cm 0,104-1,97 low -medium<br />

RD1-30cm 1,38-11,85 medium to high<br />

This study: RD1-100cm 1,18-4,67 medium<br />

Red soils RD1-400cm 2,71-12,14 medium to high<br />

Variations <strong>of</strong> values <strong>of</strong> mv, cv <strong>an</strong>d K with increased loading during oedometer consolidation<br />

tests have also been investigated by plotting them mid-way between applied pressure (log<br />

scale) values (Figs. 7.19, 7.20, 7.21 <strong>an</strong>d 7.22), since they relate to increase from one loading<br />

stage to the next.<br />

7,000E-09<br />

K/ log p curves: black <strong>clays</strong> & red soils<br />

6,000E-09<br />

Permeability K (m/s)<br />

5,000E-09<br />

4,000E-09<br />

3,000E-09<br />

2,000E-09<br />

1,000E-09<br />

0,000E+00<br />

1 10 100 1000 10000<br />

Log pressure (kPa)<br />

SA2-70cm (black <strong>clays</strong>) SA37-50cm (black <strong>clays</strong>) SB1-70cm (black <strong>clays</strong>)<br />

SB42-50cm (black <strong>clays</strong>) SC17-50cm (black <strong>clays</strong>) SC29-50cm (black <strong>clays</strong>)<br />

RD1-400cm (red soils)<br />

Figure 7.19. Permeability/ log pressure (K/log p) curves for black <strong>clays</strong> <strong>an</strong>d red soils.<br />

In Fig. (7.19), the K/ log p curves <strong>of</strong> the red soils generally plot higher up the diagram th<strong>an</strong><br />

those <strong>of</strong> black <strong>clays</strong>, <strong>an</strong>d this especially at the initial stages <strong>of</strong> loading. This serves to indicate<br />

the generally higher permeability <strong>of</strong> red soils relative to the black <strong>clays</strong>. However, the<br />

permeability <strong>of</strong> red soils is also observed to decrease with increased depth (Fig. 7.20), <strong>an</strong>d this


131<br />

is most probably due to the tendency <strong>of</strong> deeper soil horizons to be relatively more<br />

preconsolidated by increased overburden material.<br />

K/ log p curves: red soils<br />

Permeability K (m/s)<br />

4,000E-08<br />

3,000E-08<br />

2,000E-08<br />

1,000E-08<br />

0,000E+00<br />

1 10 100 1000 10000<br />

Log pressure (kPa)<br />

RD1-30cm<br />

RD1-100cm<br />

RD1-400cm<br />

Figure 7.20. Permeability/ log pressure (K/log p) curves for red soils.<br />

In Fig. (7.21), mv/ log p curves for red soils also generally plot higher th<strong>an</strong> those for black<br />

<strong>clays</strong>. This once again signifies a generally higher compressibility per unit thickness <strong>of</strong> the red<br />

soils relative to black <strong>clays</strong>. However, the compressibility <strong>of</strong> the red soils alone tends to<br />

decrease with increased depth; <strong>an</strong>d this may once again be a result <strong>of</strong> preconsolidation <strong>an</strong>d/ or<br />

overconsolidation tendencies <strong>of</strong> deeper soil horizons.<br />

mv/ log p curves: black <strong>clays</strong> & red soils<br />

3,500<br />

3,000<br />

2,500<br />

mv (m²/MN)<br />

2,000<br />

1,500<br />

1,000<br />

0,500<br />

0,000<br />

1 10 100 1000 10000<br />

Log pressure (kPa)<br />

SA2-70cm (black <strong>clays</strong>) SA37-50cm (black <strong>clays</strong>) SB1-70cm (black <strong>clays</strong>)<br />

SB42-50cm (black <strong>clays</strong>) SC17-50cm (black <strong>clays</strong>) SC29-50cm (black <strong>clays</strong>)<br />

RD1-100cm (red soils)<br />

RD1-400cm (red soils)<br />

Figure 7.21. Coefficient <strong>of</strong> volume compressibility/ log pressure (mv/log p) curves for black<br />

<strong>clays</strong> <strong>an</strong>d red soils.


132<br />

In Fig. (7.22), cv/ log p curves for red soils once again plot above those <strong>of</strong> the black <strong>clays</strong>.<br />

This implies that red soils exhibit higher rates <strong>of</strong> consolidation-settlement th<strong>an</strong> black <strong>clays</strong>.<br />

The difference, once again, could be due to the higher voids ratios <strong>an</strong>d/ or porosity (Fig. 7.18)<br />

<strong>an</strong>d higher permeability (Fig. 7.19) <strong>of</strong> the red soils relative to the black <strong>clays</strong>.<br />

18,00<br />

cv/ log p curves: black <strong>clays</strong> & red soils<br />

16,00<br />

14,00<br />

12,00<br />

cv (m²/year)<br />

10,00<br />

8,00<br />

6,00<br />

4,00<br />

2,00<br />

0,00<br />

1 10 100 1000 10000<br />

Log pressure (kPa)<br />

SA2-70cm (black <strong>clays</strong>) SA37-50cm (black <strong>clays</strong>) SB1-70cm (black <strong>clays</strong>)<br />

SB42-50cm (black <strong>clays</strong>) SC17-50cm (black <strong>clays</strong>) SC29-50cm (black <strong>clays</strong>)<br />

RD1-30cm (red soils)<br />

RD1-400cm (red soils)<br />

Figure 7.22. Coefficient <strong>of</strong> consolidation/ log pressure (cv/log p) curves for black<br />

<strong>clays</strong> <strong>an</strong>d red soils.<br />

Results <strong>of</strong> compression index, cc, obtained for the soils in this study are presented in Table<br />

(7.15b) together with corresponding values <strong>of</strong> liquid limit <strong>an</strong>d plasticity index. Based on<br />

compression index (Lambe <strong>an</strong>d Whitm<strong>an</strong>, 1979), clay soils could be approximately grouped<br />

into those <strong>of</strong> low compressibility, medium to high compressibility, high compressibility <strong>an</strong>d<br />

very high compressibility; as shown in Table (7.15a).<br />

Table 7.15a. Compression indices, Cc, <strong>an</strong>d compressibility classification<br />

<strong>of</strong> clay soils.<br />

Compression index Cc<br />

Compressibility class<br />

2,6 Very high


133<br />

The results <strong>of</strong> correlation show that liquid limit, LL, would be a better estimator <strong>of</strong> the<br />

compression index <strong>an</strong>d compressibility <strong>of</strong> soils th<strong>an</strong> the plasticity index, PI, (Table 7.15b).<br />

The strength <strong>of</strong> correlation between cc <strong>an</strong>d LL is very high (R = -0,85) for the black <strong>clays</strong>; but<br />

is reduced to moderate (R = -0,58) for combined black <strong>clays</strong> <strong>an</strong>d red soils (Figs. 7.23 <strong>an</strong>d<br />

7.24). The relationships between compression index <strong>an</strong>d liquid limit take the form <strong>of</strong><br />

cc = 0,0099(122-LL) (7.47)<br />

for black <strong>clays</strong>, <strong>an</strong>d<br />

cc = 0,0016(308-LL) (7.48)<br />

for both black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Table 7.15b. Results <strong>of</strong> compression indices correlated with Atterberg limits.<br />

Sample No. Load<br />

(kPa) Wn (%) LL (%) PI (%) Cc Approx. Compressibility<br />

SA2-70cm 100-2400 34 90 54 0,3 medium-high<br />

SA37-50cm 25-1600 22 83 44 0,4 medium-high<br />

SB1-70cm 50-2400 30 90 51 0,29 medium-high<br />

SB42-50cm 100-2400 21 91 47 0,34 medium-high<br />

SC17-50cm 100-2400 22 81 48 0,44 medium-high<br />

SC29-50cm 100-2400 20 80 43 0,38 medium-high<br />

RD1-30cm 12,0-2400 39 49 18 0,42 medium-high<br />

RD1-100cm 12,0-2400 24 51 21 0,4 medium-high<br />

RD1-400cm 12,0-2400 25 48 18 0,39 medium-high<br />

Correlations (R) Black & red soils -0,58 -0,55<br />

Correlations (R ) Black <strong>clays</strong> only -0,85 -0,66<br />

Results <strong>of</strong> correlations <strong>an</strong>d <strong>an</strong>alysis show that cc values generally decrease with increasing<br />

liquid limit <strong>an</strong>d plasticity index <strong>of</strong> soils ( Table 7.15b; Figs. 7.23 & 7.24). This serves to<br />

suggest that compressibility <strong>of</strong> soils generally decreases with increased plasticity, <strong>an</strong>d vice<br />

versa.


134<br />

Black <strong>clays</strong><br />

0,45<br />

0,43<br />

n = 6<br />

Compression index Cc<br />

0,41<br />

0,39<br />

0,37<br />

0,35<br />

0,33<br />

0,31<br />

0,29<br />

Cc = 0,0099(122-LL)<br />

R 2 = 0,73<br />

Cc vs LL<br />

0,27<br />

0,25<br />

75 80 85 90 95<br />

Liquid limit LL (%)<br />

Figure 7.23. Correlation between laboratory determined liquid limits <strong>an</strong>d<br />

compression indices for black <strong>clays</strong>.<br />

Black <strong>clays</strong> <strong>an</strong>d red soils<br />

Compression index Cc<br />

0,5<br />

0,4<br />

0,3<br />

0,2<br />

0,1<br />

0<br />

n = 9<br />

Cc = 0,0016(308-LL)<br />

R 2 = 0,34<br />

Cc vs LL<br />

0 20 40 60 80 100<br />

Liquid limit, LL (%)<br />

Figure 7.24. Correlation between laboratory determined liquid limits <strong>an</strong>d<br />

compression indices for black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Calculated values <strong>of</strong> compression index for black <strong>clays</strong>, cc(b), as well as for combined black<br />

<strong>clays</strong> <strong>an</strong>d red soils, cc(br), as derived from laboratory measured liquid limits using the<br />

relationships <strong>of</strong> Equations (7.47 & 7.48) are presented in Table (7.16); alongside laboratory<br />

determined compression indices, cc. The strength <strong>of</strong> correlation between so calculated <strong>an</strong>d<br />

laboratory measured compression indices are strong (R = 0,85) for the black <strong>clays</strong> alone; <strong>an</strong>d<br />

only moderate (R = 0,58) for combined black <strong>clays</strong> <strong>an</strong>d red soils. However, values <strong>of</strong><br />

compression indices calculated from laboratory liquid limits using Skempton´s relationship,<br />

i.e. ( Cc(sk) = 0,009(LL-10) are generally overestimated with respect to laboratory determined<br />

compression indices for black <strong>clays</strong> (Table 7.16). The overestimation could be accounted for


135<br />

by differences in lithology, mineralogy <strong>an</strong>d/ or org<strong>an</strong>ic matter content between <strong>tropical</strong> soils<br />

as encountered in this study, <strong>an</strong>d those <strong>of</strong> temperate climates on which Skempton´s <strong>an</strong>alysis<br />

<strong>an</strong>d derivation were based. This leaves the new relationships developed in this study as better<br />

<strong>an</strong>d preferred estimators <strong>of</strong> compressibility characteristics <strong>of</strong> soils using laboratory measured<br />

liquid limits.<br />

Table 7.16. Correlation: calculated <strong>an</strong>d laboratory derived compression indices.<br />

Sample No. Load kPa Wn (%) LL (%) PI (%) Cc Cc(b) Cc(br) Cc(Sk)<br />

SA2-70cm 100-2400 34 90 54 0,30 0,32 0,35 0,72<br />

SA37-50cm 25-1600 22 83 44 0,40 0,39 0,36 0,66<br />

SB1-70cm 50-2400 30 90 51 0,29 0,32 0,35 0,72<br />

SB42-50cm 100-2400 21 91 47 0,34 0,31 0,35 0,73<br />

SC17-50cm 100-2400 22 81 48 0,44 0,41 0,36 0,64<br />

SC29-50cm 100-2400 20 80 43 0,38 0,42 0,36 0,63<br />

RD1-30cm 12,0-2400 39 49 18 0,42 0,41 0,35<br />

RD1-100cm 12,0-2400 24 51 21 0,40 0,41 0,37<br />

RD1-400cm 12,0-2400 25 48 18 0,39 0,42 0,34<br />

Correlations (R ) Black <strong>clays</strong> only 0,85<br />

Correlations (R ) Black <strong>clays</strong> & red soils 0,58<br />

Future works could attempt to perform correlations between laboratory determined<br />

compression indices <strong>an</strong>d index properties, based on a larger amount <strong>of</strong> data th<strong>an</strong> is currently<br />

available in this study. This would serve to indicate possible improvements in the strength <strong>an</strong>d<br />

reliability <strong>of</strong> correlations, as well as assessment <strong>of</strong> compressibility characteristics <strong>of</strong> the soils<br />

based on laboratory measured index properties.<br />

So far, it c<strong>an</strong> be safely stated that laboratory determined liquid limits could be used to<br />

approximately estimate the compression indices <strong>an</strong>d , therefore, assess compressibility<br />

characteristics <strong>of</strong> the black <strong>clays</strong> <strong>an</strong>d red soils found.<br />

Values <strong>of</strong> secondary compression ,cα, obtained for black <strong>clays</strong> <strong>an</strong>d red soils are given in Table<br />

(7.17), <strong>an</strong>d have been used to classify the soils into various groups. Typical values <strong>of</strong> cα for<br />

other types <strong>of</strong> soil (after Lambe <strong>an</strong>d Whitm<strong>an</strong>, 1979) are also included for comparison<br />

purposes. The black <strong>clays</strong> fall in the class <strong>of</strong> normally consolidated <strong>clays</strong> while the red soils<br />

are classified as normally consolidated to slightly overconsolidated.


136<br />

Table 7.17. Classification <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils based on coefficient <strong>of</strong> secondary<br />

compression.<br />

Coeff. <strong>of</strong><br />

secondary<br />

compression Average<br />

Category Sample No. cα value cα Clay type<br />

Typical examples < 0,001 Overconsolidated <strong>clays</strong><br />

<strong>of</strong> <strong>clays</strong> 0,005-0,02 Normally consolidated <strong>clays</strong><br />

(after Lambe & 0,03 <strong>an</strong>d over Very plastic <strong>clays</strong> <strong>an</strong>d/ or<br />

Whitm<strong>an</strong> –1979)<br />

org<strong>an</strong>ic <strong>clays</strong><br />

SA2-70cm 0,0002-0,011 0,006 Normally consolidated <strong>clays</strong><br />

This study: SA37-50cm 0,003-0,017 0,009 Normally consolidated <strong>clays</strong><br />

Black <strong>clays</strong> SB1-70cm 0,0002-0,012 0,005 Normally consolidated <strong>clays</strong><br />

SB42-50cm 0,005-0,024 0,014 Normally consolidated <strong>clays</strong><br />

SC17-50cm 0,006-0,02 0,012 Normally consolidated <strong>clays</strong><br />

SC29-50cm 0,003-0,014 0,008 Normally consolidated <strong>clays</strong><br />

RD1-30cm 0,002-0,004 0,003 Slightly overconsolidated <strong>clays</strong><br />

This study: RD1-100cm 0,003-0,008 0,006 Normally consolidated <strong>clays</strong><br />

Red soils RD1-400cm 0,002-0,007 0,004 Slightly overconsolidated <strong>clays</strong><br />

Results <strong>of</strong> swell index, Cs, obtained for black <strong>clays</strong> <strong>an</strong>d red soils during unloading phase <strong>of</strong><br />

oedometer tests are summarised in Table (7.18). The black <strong>clays</strong> generally exhibit higher<br />

swelling capabilities <strong>an</strong>d/ or exp<strong>an</strong>sion tendencies (Cs = 0,03-0,075; me<strong>an</strong> value = 0,054) on<br />

unloading th<strong>an</strong> the red soils (Cs = 0,02-0,03; me<strong>an</strong> value = 0,027). Correlations between swell<br />

indices <strong>an</strong>d index properties (liquid limit, plasticity index, free swell) are generally poor<br />

(Figures 7.25, 7.26 & 7.27). This implies that uncertainties would be encountered in trying to<br />

assess <strong>an</strong>d/ or predict soils swelling <strong>an</strong>d/ or exp<strong>an</strong>sion tendencies based on results <strong>of</strong> index<br />

tests performed on disturbed <strong>an</strong>d fractioned samples.<br />

Table 7.18. Results <strong>of</strong> swell indices for black <strong>clays</strong> <strong>an</strong>d red soils correlated with Atterberg<br />

limits <strong>an</strong>d free swell.<br />

Sample No. Load kPa Wn (%) LL (%) PI (%) FS (%) Cs<br />

SA2-70cm 100-2400 34 90 54 170 0,07<br />

SA37-50cm 25-1600 22 83 44 115 0,075<br />

SB1-70cm 50-2400 30 90 51 145 0,043<br />

SB42-50cm 100-2400 21 91 47 145 0,03<br />

SC17-50cm 100-2400 22 81 48 110 0,04<br />

SC29-50cm 100-2400 20 80 43 110 0,065<br />

RD1-30cm 12,0-2400 39 49 18 20 0,03<br />

RD1-100cm 12,0-2400 24 51 21 15 0,03<br />

RD1-400cm 12,0-2400 25 48 18 15 0,02<br />

Correlations (R ) Black & red soils 0,61 0,64 0,63<br />

Correlations (R ) Black <strong>clays</strong> only -0,34 -0,14 -0,06<br />

Diagramatically, the generally weak correlations are reflected in terms <strong>of</strong> plot points which<br />

tend to be more spread out (Figures 7.25, 7.26 & 7.27).


137<br />

Black <strong>clays</strong> <strong>an</strong>d red soils<br />

Swell index Cs<br />

0,08<br />

0,07<br />

0,06<br />

0,05<br />

0,04<br />

0,03<br />

0,02<br />

0,01<br />

0<br />

n = 9<br />

40 50 60 70 80 90 100<br />

Liquid limit LL (%)<br />

LL vs Cs<br />

Figure 7.25. Correlation between swell index <strong>an</strong>d liquid limit for black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Black <strong>clays</strong> <strong>an</strong>d red soils<br />

Swell index Cs<br />

0,08<br />

0,07<br />

0,06<br />

0,05<br />

0,04<br />

0,03<br />

0,02<br />

0,01<br />

0<br />

n = 9<br />

10 20 30 40 50 60<br />

PI vs Cs<br />

Plasticity index PI (%)<br />

Figure 7.26. Correlation between swell index <strong>an</strong>d plasticity index for black<br />

<strong>clays</strong> <strong>an</strong>d red soils.<br />

Black <strong>clays</strong> <strong>an</strong>d red soils<br />

Swell index Cs<br />

0,08<br />

0,07<br />

0,06<br />

0,05<br />

0,04<br />

0,03<br />

0,02<br />

0,01<br />

0<br />

n = 9<br />

0 50 100 150 200<br />

Free swell FS (%)<br />

FS vs Cs<br />

Figure 7.27. Correlation between swell index <strong>an</strong>d free swell for black <strong>clays</strong><br />

<strong>an</strong>d red soils.


138<br />

7.4.7 Swelling pressure <strong>an</strong>d percentage swelling<br />

Black <strong>clays</strong> have been shown to exhibit considerable swelling capability, i.e. free swell <strong>of</strong><br />

between 100 <strong>an</strong>d 200% when allowed free access to water from a dry state. On the other h<strong>an</strong>d,<br />

red soils exhibit a limited swelling capability when saturated with water (free swell= 15-<br />

20%); <strong>an</strong>d as a result, no swelling pressure determination was performed on these soils.<br />

7.4.7.1 Swelling pressure test<br />

Swelling pressure, SP, <strong>of</strong> <strong>clays</strong> is the pressure required to prevent swelling by constraining the<br />

<strong>clays</strong> to maintain const<strong>an</strong>t volume when saturated with water (Nelson <strong>an</strong>d Miller, 1992).<br />

Swelling pressures <strong>of</strong> over 1 MN/m² have been registered for some heavily overconsolidated<br />

<strong>clays</strong> (Head, 1988). In this study, swelling pressures <strong>of</strong> 49-104 kPa have been determined for<br />

the black <strong>clays</strong>.<br />

Swelling pressure determination was carried out according to a method devised by Head<br />

(1988), <strong>an</strong>d employed the British St<strong>an</strong>dard oedometer consolidation cell as described in a<br />

previous section (Plate 7.9).<br />

Plate 7.9. Swelling pressure determination <strong>of</strong> clay soils.


139<br />

Testing procedure involved cutting <strong>an</strong> undisturbed soil specimen into the consolidation ring,<br />

assembling it in the oedometer cell <strong>an</strong>d mounting the cell in the load frame <strong>of</strong> a consolidation<br />

press. Distilled water was added into the cell to saturate the soil specimen <strong>an</strong>d a stop-clock<br />

started at the same time. Ch<strong>an</strong>ges in compression gauge reading due to swelling <strong>of</strong> specimen<br />

were noted <strong>an</strong>d recorded with time; <strong>an</strong>d small weights added to the weight h<strong>an</strong>ger <strong>of</strong> the<br />

apparatus to bring the compression gauge reading back to zero or initial reading <strong>an</strong>d,<br />

therefore, prevent swelling. The compression gauge reading was observed over a period <strong>of</strong> 24<br />

hours, during which time more weights were added as necessary to maintain the reading as<br />

close as possible to zero or initial reading.<br />

120<br />

Swelling Pressure curves for black <strong>clays</strong><br />

100<br />

SA41-30cm (104 kPa)<br />

Swelling pressure (kN/m²)<br />

80<br />

60<br />

40<br />

SA2-70cm (95 kPa)<br />

SB1-70cm (49 kPa)<br />

SB42-50cm (53 kPa)<br />

SC17-50cm (70 kPa)<br />

20<br />

SC29-50cm (50 kPa)<br />

0<br />

0 10 20 30 40 50 60<br />

Square-root-time (minutes)<br />

Figure 7.28. Swelling pressure curves obtained for black <strong>clays</strong> in this study.<br />

A record <strong>of</strong> the amount <strong>of</strong> each load increment <strong>an</strong>d time from start when added was tabulated<br />

(Table 7.19), <strong>an</strong>d a graph <strong>of</strong> total pressure on specimen against square root <strong>of</strong> time plotted.<br />

Swelling pressure test curves obtained for the black <strong>clays</strong> in this study are presented in Fig<br />

(7.28). Flattening <strong>of</strong> the curves serves to indicate equilibrium having been virtually reached.<br />

Table 7.19. Results <strong>of</strong> swelling pressures for black <strong>clays</strong>, alongside corresponding index<br />

properties <strong>an</strong>d clay mineralogy (smectites).<br />

Sample No. LL (%) PL (%) PI (%) FS (%) Wn (%) Cr LI<br />

% smectites<br />

(min.)<br />

SP (kPa)<br />

SA2-70cm 90 36 54 170 34 1,04 -0,04 90,00 95<br />

SA41-30cm 89 41 47 135 23 1,39 -0,39 95,00 104<br />

SB1-70cm 90 39 51 145 30 1,19 -0,19 95,00 49<br />

SB42-50cm 91 45 47 145 21 1,50 -0,50 95,00 53<br />

SC17-50cm 81 33 48 110 22 1,22 -0,22 95 70<br />

SC29-50cm 80 37 43 110 20 1,38 -0,38 95 50<br />

Correlations 0,22 -0,11 0,41 0,35 0,34 -0,31 0,31 -0,51


140<br />

The black <strong>clays</strong> exhibit signific<strong>an</strong>t swelling pressures <strong>of</strong> 49-104 kPa (Table 7.19). However,<br />

no well defined relationship was found to occur between swelling pressure determined on<br />

undisturbed specimens on one h<strong>an</strong>d; <strong>an</strong>d index properties as well as smectite content <strong>of</strong> the<br />

<strong>clays</strong>, on the other. Attempted correlations are generally poor <strong>an</strong>d <strong>of</strong> low strength (R< 0,6) as<br />

shown in Table (7.19); <strong>an</strong>d evidenced by more spread out plot points on scatter diagrams<br />

(Figures 7.29, 7.30, 7.31, 7.32 & 7.33). This shows the unreliability with which swelling<br />

characteristics <strong>of</strong> the <strong>clays</strong> could be estimated <strong>an</strong>d / or predicted from results <strong>of</strong> index tests.<br />

Black <strong>clays</strong><br />

Black <strong>clays</strong><br />

Swelling pressure SP (kPa)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n=7<br />

LL vs SP<br />

Swelling pressure SP (kPa)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n=7<br />

PI vs SP<br />

0<br />

0<br />

75 80 85 90 95<br />

30 40 50 60<br />

Liquid limit LL (%)<br />

Plasticity index PI (%)<br />

Figure 7.29. Correlation between LL <strong>an</strong>d SP.<br />

Figure 7.30. Correlation between PI <strong>an</strong>d SP.<br />

Once again therefore, serious uncertainties would still be encountered in attempting to assess<br />

the possible magnitude <strong>of</strong> swelling pressures <strong>of</strong> the clay soils in situ based on results <strong>of</strong> index<br />

tests performed on disturbed <strong>an</strong>d fractioned soil samples.<br />

Black <strong>clays</strong><br />

Black <strong>clays</strong><br />

120<br />

120<br />

Swelling pressure SP (kPa)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n = 7<br />

Cr vs SP<br />

Swelling pressure SP (kPa)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n = 7<br />

FS vs SP<br />

0<br />

0,80 1,00 1,20 1,40 1,60<br />

Relative consistency Cr<br />

0<br />

80 130 180<br />

Free swell FS (%)<br />

Figure 7.31. Correlation between Cr <strong>an</strong>d SP.<br />

Figure 7.32. Correlation between FS <strong>an</strong>d SP.


141<br />

Black <strong>clays</strong><br />

Swelling pressure SP (kPa)<br />

120<br />

n = 7<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

10 15 20 25 30 35 40<br />

Natural moisture Wn (%)<br />

Wn vs SP<br />

Figure 7.33. Correlation between Wn <strong>an</strong>d SP.<br />

7.4.7.2 Swelling test<br />

Swelling is the increase in volume <strong>of</strong> a soil due to absorption <strong>of</strong> water within the voids when<br />

the applied stress is reduced. It is therefore the reverse <strong>an</strong>d opposite process <strong>of</strong> consolidation,<br />

<strong>an</strong>d occurs when overconsolidated <strong>clays</strong> are allowed free access to water (Nelson <strong>an</strong>d Miller,<br />

1992). In consolidation tests, swelling is represented by the unloading portion <strong>of</strong> the e/ log p<br />

curve (Fig. 7.18 ). The mech<strong>an</strong>ism <strong>of</strong> swelling is explained by <strong>an</strong> overconsolidated soil<br />

possessing high suction tensions within its skeleton on unloading, thereby drawing water into<br />

the voids. The resulting increase in volume <strong>of</strong> voids causes the soil to swell <strong>an</strong>d eventually<br />

disintegrate.<br />

Testing procedure involved use <strong>of</strong> same apparatus as in the st<strong>an</strong>dard oedometer consolidation<br />

test. However, the height <strong>of</strong> soil specimen was 3 mm less th<strong>an</strong> the height <strong>of</strong> consolidation<br />

ring; so that the specimen was laterally confined at all times during swelling. The space in the<br />

ring above test specimen was taken up by a flat metal disc <strong>of</strong> 3 mm thickness <strong>an</strong>d diameter<br />

about 1 mm less th<strong>an</strong> the internal diameter <strong>of</strong> ring. The specimen in the ring was weighed,<br />

assembled in oedometer cell <strong>an</strong>d mounted in the load frame <strong>of</strong> consolidation press. Initial<br />

height <strong>of</strong> specimen was given by height <strong>of</strong> ring less the average disc thickness.<br />

Swelling pressure, SP, was determined as described above. After establishment <strong>of</strong><br />

equilibrium, the swelling test was performed by unloading the specimen in stages from the<br />

swelling pressure, SP, as the starting point; <strong>an</strong>d by following <strong>an</strong> unloading sequence <strong>of</strong><br />

halving the loads at each successive decrement, i.e. 1/2SP, 1/4SP, 1/8SP, <strong>an</strong>d so on down to<br />

the smallest load required. Each unloading stage was maintained until equilibrium had been<br />

reached <strong>an</strong>d no further ch<strong>an</strong>ge in compression gauge reading was observed. During<br />

unloading, cumulative amount <strong>of</strong> swell (mm) undergone by the specimen for successive<br />

decrements <strong>of</strong> load was recorded by noting the ch<strong>an</strong>ge in compression gauge reading from its<br />

initial position.<br />

The cumulative swell (S mm) under each load decrement was recorded <strong>an</strong>d expressed as a<br />

percentage <strong>of</strong> the ultimate amount <strong>of</strong> swell (Smax mm) which was registered with specimen<br />

under zero load. Results <strong>of</strong> cumulative percentage swelling (S %) recorded against


142<br />

corresponding load decrements (P kPa) <strong>an</strong>d as obtained for black <strong>clays</strong> in this study are<br />

presented in Appendix (B). A summary <strong>of</strong> results <strong>of</strong> percentage swelling under various<br />

loading for tested samples is given in Table 7.20. The variation <strong>of</strong> percentage swelling (S)<br />

with pressure loads (P) is presented in Fig. (7.34). The variation is generally similar <strong>an</strong>d close<br />

for all tested samples, <strong>an</strong>d this once again reflects on the general homogeneity <strong>of</strong> black <strong>clays</strong>.<br />

Table 7.20. Percentage swelling <strong>of</strong> black <strong>clays</strong> under various load decrements (S% in relation<br />

to ultimate amount <strong>of</strong> swelling Smax).<br />

Sample No. S/Smax S (%) P (kPa) P/SP P (%)<br />

0,01 1 82,46 1,00 100<br />

0,1 10 41,23 0,50 50<br />

SA2/70cm 0,24 24 19,99 0,24 24<br />

0,31 31 10,00 0,12 12<br />

0,52 52 5,00 0,06 6<br />

0,7 70 2,50 0,03 3<br />

1 100 0,01 0,00 0<br />

0,01 1 37,48 1,00 100<br />

0,13 13 18,74 0,50 50<br />

SB1/70cm 0,3 30 10,00 0,27 27<br />

0,46 46 5,00 0,13 13<br />

0,66 66 2,50 0,07 7<br />

1 100 0,01 0,00 0<br />

0,01 1 36,23 1,00 100<br />

0,14 14 17,49 0,48 48<br />

SC25/50cm 0,34 34 8,75 0,24 24<br />

0,55 55 5,00 0,14 14<br />

0,73 73 2,50 0,07 7<br />

1 100 0,01 0,00 0<br />

0,01 1 44,98 1,00 100<br />

0,12 12 22,49 0,50 50<br />

SC29/50cm 0,3 30 11,24 0,25 25<br />

0,54 54 5,00 0,11 11<br />

0,72 72 2,50 0,06 6<br />

1 100 0,01 0,00 0


143<br />

Black <strong>clays</strong>: P (kPa) vs S (%)<br />

Pressure load P (kPa)<br />

90,00<br />

80,00<br />

70,00<br />

60,00<br />

50,00<br />

40,00<br />

30,00<br />

20,00<br />

10,00<br />

0,00<br />

0 20 40 60 80 100 120<br />

Swelling S (%)<br />

SA2/70cm<br />

SB1/70cm<br />

SC25/50cm<br />

SC29/50cm<br />

Figure 7.34. Variation <strong>of</strong> percentage swelling <strong>of</strong> black <strong>clays</strong> with loading pressures.<br />

A classification <strong>of</strong> the relative degree <strong>of</strong> swelling achieved under various loading pressure<br />

decrements for each <strong>of</strong> the soil samples tested is given in Table (7.21). As a result, a general<br />

classification for the black <strong>clays</strong> as a whole, has been derived <strong>an</strong>d presented in Table (7.22). It<br />

is observed that black <strong>clays</strong> would exhibit high percentage swelling when externally loaded to<br />

below 5 kPa; <strong>an</strong>d very low percentage swelling when loaded to over 80 kPa. This implies that<br />

external <strong>engineering</strong> structures imposing loads <strong>of</strong> up to 5 kPa <strong>an</strong>d less would still give way to<br />

maximum swelling <strong>of</strong> the <strong>clays</strong> on wetting, with a corresponding maximum destabilisation <strong>of</strong><br />

the same structures. However, the structures would be relatively more stable when designed to<br />

impose loads <strong>of</strong> over 80 kPa.<br />

Table 7.21. Relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> under various load<br />

decrements.<br />

Swelling degree<br />

Sample No. S (%) P (kPa) (after author, 2003)<br />

SA2/70cm 100-52 0-4,96 high<br />

52-31 4,96-10 moderate<br />

31-0 10,0-82,46 low<br />

SB1/70cm 100-85 0-1,18<br />

high<br />

85-30 1,18-10 moderate<br />

30-0 10,0-37,48 low<br />

SC25/50cm 100-75 0-1,70 high<br />

75-34 1,70-8,75<br />

moderate<br />

34-0 8,75-36,23 low<br />

SC29/50cm 100-85 0-1,58 high<br />

85-30 1,58-11,24 moderate<br />

30-0 11,24-44,98 low


144<br />

Table 7.22. Derived classification <strong>of</strong> relative degree <strong>of</strong> swelling <strong>of</strong> black<br />

<strong>clays</strong> based on percentage swelling (after author, 2003).<br />

P (kPa) S (%) Swelling degree<br />

0-2,5 100-75 high<br />

2,5-10 75-30,0 moderate<br />

10-50,0 30-10,0 low<br />

> 50,0 < 10 very low<br />

A correlation between percentage swelling <strong>an</strong>d corresponding load decrements (expressed as<br />

percentage <strong>of</strong> respective swelling pressures) for the individual soil samples is presented<br />

diagrammatically in Fig. (7.35). The variation is generally logarithmic with a very strong<br />

correlation (R = 0,99 – 1,0).<br />

Individual samples: S (%) vs P (%)<br />

120<br />

Loading P (%)<br />

100<br />

80<br />

60<br />

40<br />

R 2 = 0,98<br />

R 2 = 0,99<br />

R 2 = 1,0<br />

SA2-70cm<br />

SB1-70cm<br />

SC25-50cm<br />

SC29-50cm<br />

Logarithmic SA2-70cm<br />

Logarithmic SB1-70cm<br />

20<br />

0<br />

R 2 = 1,0<br />

0 50 100 150<br />

Logarithmic SC25-50cm<br />

Logarithmic SC29-50cm<br />

Swelling S (%)<br />

Figure 7.35. Correlation between percentage swelling <strong>an</strong>d percent loading pressure <strong>of</strong><br />

individual black clay samples.<br />

Results <strong>of</strong> <strong>an</strong> overall correlation combining results <strong>of</strong> all samples tested are presented Fig.<br />

(7.36). The relationship between percentage swelling <strong>an</strong>d extent <strong>of</strong> loading is once again<br />

logarithmic (Fig. 7.36a) with a very strong correlation (R = 0,995), <strong>an</strong>d could be summarised<br />

by <strong>an</strong> overall equation, i.e.<br />

P = -22,3Ln(S) + 100,8 (7.49)<br />

where P = extent <strong>of</strong> loading (%)<br />

S = cumulative swelling (%)


145<br />

The same relationship is represented by a straight-line graph by plotting S% on a logarithmic<br />

scale (Fig. 7.36b).<br />

The above relationship could be used as a guide to estimate the amount <strong>of</strong> external loading (P<br />

kPa) necessary to produce <strong>an</strong> allowed amount <strong>of</strong> swelling (S mm) or percentage swelling<br />

(S%) <strong>of</strong> the <strong>clays</strong>, as may be required for <strong>engineering</strong> design purposes, if the swelling<br />

pressure (SP), as well as ultimate swelling (Smax) under zero loading, are already known. For<br />

inst<strong>an</strong>ce, limited swelling <strong>of</strong> the <strong>clays</strong> ( i.e. S/ Smax < 0,01or S% < 1) would only be realised<br />

in practice with external loads (P ) close to the swelling pressure (SP), i.e. P% (or P/SP *100)<br />

is close or equal to 100. This could also be confirmed by the above established relationship<br />

by substituting for S/Smax = 0,01 (or S% = 1), i.e.<br />

P% = -22,3Ln(1) + 100,8 = 0 + 100,8, or<br />

P% = (P/SP)*100 = 100,8<br />

On the other h<strong>an</strong>d, zero external loading conditions would be expected to allow for maximum<br />

swelling <strong>of</strong> the <strong>clays</strong>, i.e. S% = 100, in normal practice. However, substituting P% (i.e. P = 0<br />

kPa) in the above relationship gives <strong>an</strong> underestimated value <strong>of</strong> S% = 92, i.e.<br />

0 = -22,3Ln(S) + 100,8 or Ln(S) = 100,8/22,3<br />

so that S = e^4,52 = 92%<br />

The underestimation <strong>of</strong> percentage swelling from known loading pressures could be explained<br />

by the fact that, test specimens <strong>of</strong> <strong>clays</strong> underwent some amount <strong>of</strong> plastic deformation (or<br />

consolidation) when initially loaded to their swelling pressure values, <strong>an</strong>d could not therefore<br />

recover their original sizes on complete unloading. In this case, the extent <strong>of</strong> plastic<br />

deformation is equivalent to S% = 8 (i.e., 100-92), on the average. A suggested remedy <strong>of</strong> this<br />

shortcoming would be to determine the ultimate amount <strong>of</strong> swelling (Smax) under zero<br />

loading on separate test specimens from those used in the swelling test. Cumulative swelling<br />

values obtained during the test would then be expressed as a percentage <strong>of</strong> this separately<br />

determined maximun swelling value <strong>of</strong> Smax.<br />

Future works on swelling tests involving larger clay sample sizes <strong>an</strong>d data could yield more<br />

closely approximating logarithmic relationships with relatively stronger correlations.


146<br />

Combined sample results: S (%) vs P (%)<br />

Loading P (%)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

n = 25<br />

P = -22,3Ln(S) + 100,8<br />

R 2 = 0,99<br />

Combined sample<br />

results<br />

Logarithmic<br />

20<br />

0<br />

0 50 100 150<br />

Swelling S (%)<br />

Figure 7.36a. Correlation between percentage swelling <strong>an</strong>d extent <strong>of</strong> loading using<br />

combined sample results <strong>of</strong> black <strong>clays</strong>.<br />

Combined sample results: S (%) vs P (%)<br />

120<br />

n = 25<br />

100<br />

P= -22,3Ln(S) + 100,8<br />

Loading P (%)<br />

80<br />

60<br />

R 2 = 0,99<br />

Combined<br />

sample results<br />

40<br />

20<br />

0<br />

1 10 100<br />

Swelling S (%) on Log scale<br />

Figure 7.36b. Correlation between percentage swelling (log scale) <strong>an</strong>d extent <strong>of</strong> loading using<br />

combined sample results <strong>of</strong> black <strong>clays</strong>.


147<br />

Alternatively, percentage swelling <strong>of</strong> <strong>clays</strong> could be derived by expressing the swelling (S<br />

mm) which occurs under <strong>an</strong>y load decrement (P kPa) as a percentage <strong>of</strong> the initial thickness<br />

(Ho mm) <strong>of</strong> test specimen. Complete results <strong>of</strong> percentage swelling calculated for tested black<br />

clay samples using this second method are given in Appendix C. A more summarised<br />

representation <strong>of</strong> the results for the samples is given in Table (7.23). A derived classification<br />

<strong>of</strong> relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> based on this new method is also given in Table<br />

(7.24).<br />

Table 7.23. Percentage swelling <strong>of</strong> black <strong>clays</strong> under various load decrements (S% is in<br />

relation to initial specimen height, Ho, i.e. 2 nd method).<br />

Sample No. P (kPa) SP = Pmax (kPa) P (%) S (mm) Smax (mm) S (%) Smax(%)<br />

82,46 100 0,000 0,00<br />

41,23 50 0,106 0,96<br />

SA2-70cm 19,99 24 0,253 2,30<br />

10,00 82,46 12 0,321 1,036 2,92 9,42<br />

5,00 6 0,539 4,90<br />

2,50 3 0,725 6,59<br />

0,00 0 1,036 9,42<br />

37,48 100 0,000 0,00<br />

18,74 50 0,054 0,49<br />

SB1-70cm 10,00 37,48 27 0,126 0,425 1,15 3,86<br />

5,00 13 0,197 1,79<br />

2,50 7 0,279 2,54<br />

0,00 0 0,425 3,86<br />

36,23 100 0 0,00<br />

17,49 48 0,066 0,60<br />

SC25/50cm 8,75 36,23 24 0,16 0,475 1,45 4,32<br />

5,00 14 0,26 2,36<br />

2,50 7 0,347 3,15<br />

0,00 0 0,475 4,32<br />

44,98 100 0 0,00<br />

22,49 50 0,061 0,55<br />

SC29-50cm 11,24 44,98 25 0,149 0,493 1,35 4,48<br />

5,00 11 0,266 2,42<br />

2,50 6 0,356 3,24<br />

0,00 0 0,493 4,48<br />

Average 50,29 0,607 5,52<br />

Table 7.24. Classification <strong>of</strong> relative degree <strong>of</strong> swelling <strong>of</strong> black <strong>clays</strong> based<br />

on percentage swelling derived using the new/ 2 nd method (author, 2003).<br />

P (kPa) P (%) S (%) Swelling degree<br />

< 2,5 < 5 > 5 high<br />

2,5-7,5 5-15 5 – 2,5 moderate<br />

7,5-25,0 15-50 2,5- 0,5 low<br />

> 25,0 > 50 < 0,5 very low


148<br />

The variation <strong>of</strong> so calculated percentage swelling (S%) with extent <strong>of</strong> external loading P%<br />

(i.e., load decrements, P kPa, expressed as a percentage <strong>of</strong> respective swelling pressure, SP<br />

kPa) is presented diagrammatically in Fig. (7.37a). The variation is also closely logarithmic<br />

with a very strong correlation (R = 0,98), i.e.<br />

P = -15,86Ln(S) + 29,84 (7.50)<br />

where P = extent <strong>of</strong> loading (%)<br />

S = cumulative swelling (%), derived w.r.t initial specimen height Ho<br />

This relationship is also represented as a straight line graph in Fig. (7.37b) where percentage<br />

swelling (S%) has been plotted on a logarithmic scale.<br />

The second relationship could also be possibly used to estimate <strong>an</strong>d/ or <strong>an</strong>ticipate percentage<br />

swelling (S%) <strong>an</strong>d/ or swelling (S mm) that may arise from certain or chosen projected<br />

external loading conditions. For inst<strong>an</strong>ce, choosing zero loading conditions implies P (kPa)<br />

<strong>an</strong>d P% [i.e. (P/SP)*100] are also zero, so that from Equation (7.50),<br />

0 = -15,86Ln(S) + 29,84 or Ln(S) = 29,84/15,86 = 1,88, so that<br />

S = e^ 1,88 = 6,55%<br />

Or for initial specimen thickness Ho <strong>of</strong> 11mm,<br />

S = (6,55/100) * 11 = 0,72mm<br />

On the average, therefore, the black <strong>clays</strong> would undergo a maximum percentage swelling <strong>of</strong><br />

6,55% (or swelling <strong>of</strong> 0,72 mm relative to the initial specimen thickness Ho <strong>of</strong> 11 mm), under<br />

zero external loading conditions. Other values <strong>of</strong> percentage swelling <strong>an</strong>d/ or amount <strong>of</strong><br />

swelling that may arise from selected external loading could be similarly calculated or just<br />

read <strong>of</strong>f from the curves <strong>of</strong> Figs. (7.37a &b).<br />

The second relationship could also be possibly used for estimating the extent <strong>of</strong> external<br />

loading (P%) <strong>an</strong>d/ or external load (P kPa) necessary to give chosen or allowable percentage<br />

swelling <strong>an</strong>d/ or amount <strong>of</strong> swelling, as may be required by foundation designers. For<br />

inst<strong>an</strong>ce, a selected average percentage swelling <strong>of</strong> S% = 2 (or S = 0,22 mm for Ho = 11 mm)<br />

would be exhibited by the <strong>clays</strong> when the extent <strong>of</strong> external loading P% is such that<br />

P% = -15,86 * Ln(2) + 29,84 = -(15,86 * 0,693) + 29,84, or<br />

P% = 18,85,<br />

<strong>an</strong>d this is the same value given by the curve <strong>of</strong> Fig. (7.37) for S% = 2.<br />

For <strong>an</strong> average swelling pressure, SP = 50,29 kPa as calculated for tested samples <strong>of</strong> black<br />

<strong>clays</strong> in Table (7.23), the external loading pressure is given by<br />

P (kPa) = (P% * SP)/100 = (18,85 * 50,29)/100 or<br />

P (kPa) = 9,48


149<br />

Combined sample results: S % vs P% (2nd method)<br />

120<br />

100<br />

n = 25<br />

Loading P (%)<br />

80<br />

60<br />

40<br />

P = -15,86Ln(S) + 29,84<br />

R 2 = 0,96<br />

Combined<br />

sample results<br />

20<br />

0<br />

0,00 2,00 4,00 6,00 8,00 10,00<br />

Swelling S (%)<br />

Figure 7.37a. Correlation between percentage swelling (derived w.r.t initial specimen height<br />

Ho) <strong>an</strong>d extent <strong>of</strong> external loading for combined sample results <strong>of</strong> black <strong>clays</strong>.<br />

Combined sample results: S % vs P% (2nd method)<br />

120<br />

n = 25<br />

Loading P (%)<br />

100<br />

80<br />

60<br />

40<br />

P = -15,86Ln(S) + 29,84<br />

R 2 = 0,96<br />

Combined<br />

sample results<br />

20<br />

0<br />

0,01 0,10 1,00 10,00<br />

Swelling S (%) on log scale<br />

Figure 7.37b. Correlation between percentage swelling (derived w.r.t initial specimen height<br />

Ho) on log scale, <strong>an</strong>d extent <strong>of</strong> external loading for combined sample results <strong>of</strong> black <strong>clays</strong>.


150<br />

In comparison, black <strong>clays</strong> from field locations representing sample No. SA2-70cm alone<br />

have a swelling pressure <strong>of</strong> SP = 82,46 kPa, <strong>an</strong>d would therefore require <strong>an</strong> external load <strong>of</strong><br />

15,54 kPa to permit the same amount <strong>of</strong> percentage swelling, S% = 2, i.e.<br />

P (kPa) = (18,85 * 82,46)/100 = 15,54<br />

Other extents <strong>of</strong> external loading necessary to give certain permissible percentage swelling<br />

could be similarly calculated if the swelling pressure (SP) <strong>of</strong> the <strong>clays</strong> is already known.<br />

Tsiambaos <strong>an</strong>d Tsaligopoulos (1995) investigated swelling characteristics <strong>of</strong> exp<strong>an</strong>sive <strong>clays</strong><br />

based on examples from Greece. As a way <strong>of</strong> estimating the swelling characteristics, they<br />

suggested plotting the ratio <strong>of</strong> percentage swelling (w.r.t original specimen height, Ho) to<br />

corresponding load decrement, S (%)/P (kPa) on log scale, against the ratio <strong>of</strong> load decrement<br />

to swelling pressure, [P (kPa)/SP (kPa)] * 100%. Results <strong>of</strong> such a relationship as applied to<br />

data <strong>of</strong> black <strong>clays</strong> obtained in this study are presented in Table (7.25) <strong>an</strong>d Fig.(7.38).<br />

Table 7.25. Computed values <strong>of</strong> ratio S(%)/P <strong>an</strong>d P/SP (%) (according to Greek method)<br />

for black <strong>clays</strong> in this study.<br />

Sample No. P (kPa) SP = Pmax (kPa) Smax (mm) S (%) S (%) / P P (%) = (P/SP)*100<br />

82,46 0,00 0,01 100<br />

41,23 0,96 0,02 50<br />

SA2/70cm 19,99 82,46 1,036 2,30 0,12 24<br />

10,00 2,92 0,29 12<br />

5,00 4,90 0,98 6<br />

2,50 6,59 2,64 3<br />

37,48 0,00 0,01 100<br />

18,74 0,49 0,03 50<br />

SB1/70cm 10,00 37,48 0,425 1,15 0,12 27<br />

5,00 1,79 0,36 13<br />

2,50 2,54 1,02 7<br />

36,23 0,00 0,01 100<br />

17,49 0,60 0,03 48<br />

SC25/50cm 8,75 36,23 0,475 1,45 0,17 24<br />

5,00 2,36 0,47 14<br />

2,50 3,15 1,26 7<br />

44,98 0,00 0,01 100<br />

22,49 0,55 0,02 50<br />

SC29/50cm 11,24 44,98 0,493 1,35 0,12 25<br />

5,00 2,42 0,48 11<br />

2,50 3,24 1,30 6<br />

Average value: 50,29 0,607(5,52%)


151<br />

Black <strong>clays</strong>: Greek method<br />

P% = (P/SP) * 100<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n = 21<br />

P/SP = -17,32Ln(S/P) + 1,18<br />

R 2 = 0,85<br />

S/P vs P/SP (%)<br />

0<br />

0,00 0,50 1,00 1,50 2,00 2,50 3,00<br />

S (%)/P (kPa)<br />

Fig. 7.38a. Correlation between ratios S(%)/P <strong>an</strong>d P/SP (%), i.e. Greek method, for combined<br />

sample results <strong>of</strong> black <strong>clays</strong> ( S% derived w.r.t initial specimen height Ho).<br />

Black <strong>clays</strong>: Greek method<br />

P% = (P/SP) * 100<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

n = 21<br />

P/SP = -17,32Ln(S/P) + 1,18<br />

R 2 = 0,85<br />

S/P vs P/SP (%)<br />

0<br />

0,01 0,10 1,00 10,00<br />

S (%)/ P (kPa)<br />

Figure 7.38b. Correlation between S(%)/P on log scale <strong>an</strong>d P/SP (%) for combined sample<br />

results <strong>of</strong> black <strong>clays</strong>, based on the Greek method (S% w.r.t initial specimen height, Ho).<br />

Correlations using the Greek method resulted in the following relationship, i.e.<br />

P (%) = -17.32Ln(S/SP) + 1,18 (7.51)<br />

Where S (%) = percentage swelling in relation to initial specimen height, Ho mm.<br />

P (kPa) = load decrement<br />

SP (kPa) = swelling pressure<br />

P% = [P(kPa)/SP (kPa)] * 100


152<br />

For loading conditions equal to swelling pressure, SP, (i.e. P% or P/SP * 100 = 100%), the<br />

percentage swelling is usually expected to be zero, i.e. S% = 0. However, percentage swelling<br />

value given by the Greek method (Eq. 7.51) for these loading conditions is overestimated, i.e.<br />

by substitution,<br />

100 = -17.32Ln(S/SP) + 1,18, or<br />

S/SP = e^-5,706 = 0,0033<br />

For <strong>an</strong> average swelling pressure <strong>of</strong> SP = 50,29 kPa for black <strong>clays</strong>,<br />

S% = 0,0033* 50,29 or<br />

S% = 0,17<br />

In comparison, the second method (Eq. 7.50) developed in this study gives a more reliable<br />

estimation <strong>of</strong> S% for the same loading conditions above, i.e. S% = 0,01.<br />

On the other h<strong>an</strong>d, estimation <strong>of</strong> loading pressures necessary to realise permitted percentage<br />

swelling values using the Greek method will require solution <strong>of</strong> rather cumbersome<br />

mathematical equations <strong>of</strong> the form<br />

aP + bLn(P) + c = 0,<br />

where P = the loading pressure (kPa)<br />

a, b, c = const<strong>an</strong>ts<br />

In addition, the strength <strong>of</strong> estimating swelling characteristics by the Greek method (Eq. 7.51)<br />

is inferior (R = 0,92) relative to that <strong>of</strong> the second method (Eq. 7.50) developed in this study<br />

(R = 0,98).<br />

In conclusion, the second method (Eq. 7.50) developed in this study would be preferred <strong>an</strong>d<br />

more reliable th<strong>an</strong> the Greek one in estimating swelling characteristics <strong>of</strong> black <strong>clays</strong>.<br />

Abebe Solomon (2002), undertook a laboratory investigation <strong>of</strong> the swelling characteristics <strong>of</strong><br />

partially saturated <strong>an</strong>d remoulded highly exp<strong>an</strong>sive clay materials in Germ<strong>an</strong>y. He adopted<br />

<strong>an</strong> improved testing procedure <strong>an</strong>d apparatus in the form <strong>of</strong> a modified triaxial compression<br />

machine <strong>an</strong>d/ or cell to determine swelling pressure <strong>an</strong>d associated volume ch<strong>an</strong>ge (swelling<br />

strain) simult<strong>an</strong>eously. The obtained results could be presented in the form <strong>of</strong> curves <strong>of</strong> both<br />

swelling pressure <strong>an</strong>d moisture/ water up take variations with time, the interpretation <strong>of</strong> which<br />

serves to provide more accurate insights regarding swelling behaviour <strong>an</strong>d capability <strong>of</strong> <strong>clays</strong><br />

soils. However, his <strong>engineering</strong> computations <strong>of</strong> swelling strain rate (in terms <strong>of</strong> variations <strong>of</strong><br />

swelling stress <strong>an</strong>d moisture) are beyond the scope <strong>of</strong> the present work.<br />

7.4.8 Evaluation <strong>of</strong> results <strong>of</strong> oedometer consolidation, swelling pressure <strong>an</strong>d<br />

percentage swelling<br />

Results <strong>of</strong> consolidation parameters cv, mv, cc <strong>an</strong>d secondary compression, cα, show the red<br />

soils to be slightly overconsolidated, especially the deeper horizons (Table 7.17), so that they<br />

have in the past been subjected to a pressure greater th<strong>an</strong> the present overburden pressure.<br />

This is most probably due to the <strong>clays</strong> having been covered by deposits <strong>of</strong> soil or rock<br />

(volc<strong>an</strong>ic lava), <strong>an</strong>d which were subsequently eroded away in the course <strong>of</strong> <strong>geological</strong> time;<br />

leading to a reduction in effective pressure. Preconsolidation effects could have also been


153<br />

partly produced by effects <strong>of</strong> weathering characterised by removal <strong>of</strong> soluble bases <strong>an</strong>d<br />

minerals, <strong>an</strong>d partly due to partial drying; thereby causing a reduction in effective pressure in<br />

the soil mass. As a result, the soils tend to exhibit a stiff to hard consistency with depth.<br />

However, upper sections <strong>of</strong> soils up to 1,5 - 2m depth are generally s<strong>of</strong>t to firm, <strong>an</strong>d this is<br />

most likely due to reduced preconsolidation effects (most probably accounted for by chemical<br />

weathering <strong>an</strong>d partial drying only).<br />

In spite <strong>of</strong> overconsolidation, the red soils generally exhibit limited percentage swelling <strong>an</strong>d<br />

swelling pressure effects when allowed free access to water. This could be explained by the<br />

relatively high content <strong>of</strong> iron oxide in the soils; <strong>an</strong>d this has the effect <strong>of</strong> cementing the clay<br />

particles together, thereby counteracting, inhibiting <strong>an</strong>d/ or suppressing possible swelling <strong>an</strong>d<br />

swelling pressure effects that could arise from overconsolidation character on saturation with<br />

water (Fig. 7.18).<br />

Results <strong>of</strong> swelling tests performed on black <strong>clays</strong> show variation <strong>of</strong> percentage swelling<br />

(S%) with progressive load decrements (P kPa) to be logarithmic in relationship with a very<br />

strong correlation (R = 0,99), when the external loading is expressed as a percentage <strong>of</strong> the<br />

swelling pressure (SP kPa) <strong>an</strong>d amount <strong>of</strong> swelling (S mm) as a percentage <strong>of</strong> ultimate<br />

swelling (Smax mm) which occurs under zero external loading. The above relationship could<br />

be used as a guide to estimate the amount <strong>of</strong> external loading (P kPa) necessary to produce <strong>an</strong><br />

allowed amount <strong>of</strong> swelling (S mm) or percentage swelling (S%) <strong>of</strong> the <strong>clays</strong>, as may be<br />

required for <strong>engineering</strong> design purposes, if the swelling pressure (SP), as well as ultimate<br />

swelling (Smax) under zero loading, are already known. In general, the <strong>clays</strong> would exhibit<br />

high percentage swelling (S > 75%) when imposed loads from light <strong>engineering</strong> structures<br />

located on <strong>an</strong>d/ or within, fall below 5 kPa. On the other h<strong>an</strong>d, external imposed loads greater<br />

th<strong>an</strong> 80 kPa would me<strong>an</strong> minimal swelling <strong>of</strong> the <strong>clays</strong> (S < 10%).<br />

The swelling characteristics <strong>of</strong> black <strong>clays</strong> could also be expressed in terms <strong>of</strong> the percentage<br />

swelling which occurs with respect to initial test specimen thickness, Ho mm [i.e., S% =<br />

(S/Ho) * 100] <strong>an</strong>d extent <strong>of</strong> loading (external loading, P kPa, expressed as a percentage <strong>of</strong><br />

swelling pressure, SP kPa). This second variation <strong>an</strong>d/ or relationship is also logarithmic with<br />

a similarly very strong correlation (R = 0,98). For a known swelling pressure, SP, <strong>of</strong> the <strong>clays</strong>,<br />

this relationship may be used for the estimation <strong>of</strong> percentage swelling which could occur for<br />

a selected external loading, as well as external loading conditions necessary to give rise to<br />

certain allowable percentage swelling; with the purpose <strong>of</strong> meeting specific design<br />

requirements.<br />

In practice, preconsolidation pressure, pc, <strong>an</strong>d overconsolidation ratio, OCR (i.e OCR = pc/po;<br />

po = overburden pressure), are derived through <strong>engineering</strong> <strong>an</strong>alysis methods by constructing<br />

field consolidation curves representing soils loaded in situ from laboratory consolidation<br />

curves (Casagr<strong>an</strong>de, 1936; Leonards, 1962; Schmertm<strong>an</strong>n, 1953/54). The <strong>an</strong>alysis usually<br />

relies on there being sufficient load increments in the laboratory test to obtain three points on<br />

a straight line, as well as data on specific gravity <strong>of</strong> soil particles <strong>an</strong>d Atterberg limits for<br />

cross –checking the compression index, cc. However, no such construction <strong>of</strong> field<br />

consolidation curves was done in this study due to limitations imposed by the capacity <strong>of</strong><br />

consolidation press used.<br />

On the other h<strong>an</strong>d, results <strong>of</strong> consolidation parameters <strong>an</strong>d secondary compression values<br />

(Table 7.17) show the black <strong>clays</strong> as being normally consolidated, so that the soils have never<br />

been subjected to <strong>an</strong> effective stress greater th<strong>an</strong> the present effective overburden pressure.<br />

They are generally s<strong>of</strong>t to firm <strong>an</strong>d stiff with depth. The large percentage swelling <strong>an</strong>d


154<br />

swelling pressures exhibited by these <strong>clays</strong> when allowed access to water are therefore a<br />

result <strong>of</strong> the high content <strong>of</strong> exp<strong>an</strong>sive clay minerals (smectites) found within, rather th<strong>an</strong><br />

overconsolidation character.<br />

The black <strong>clays</strong> studied <strong>an</strong>d sampled during the dry season were observed to exhibit strong<br />

dessication cracks <strong>an</strong>d a firm to stiff, sometimes hard consistency in the field. Later laboratory<br />

studies have shown the <strong>clays</strong> to posses high swell <strong>an</strong>d exp<strong>an</strong>sion potential by exhibiting large<br />

percentage swelling <strong>an</strong>d high swelling pressures when allowed free access to water (Table<br />

7.19; Fig. 7.28). The appreciable amount <strong>of</strong> swelling <strong>an</strong>d swelling pressures which develop on<br />

saturation <strong>of</strong> the soils with water could be due partly to some form <strong>of</strong> slight seasonal<br />

overconsolidation <strong>of</strong> black <strong>clays</strong>, previously preconsolidated by effects <strong>of</strong> lowering <strong>of</strong> the<br />

ground water table <strong>an</strong>d partial drying during the hot <strong>an</strong>d dry season; <strong>an</strong>d partly to the high<br />

smectite content, as already stated.


155<br />

Chapter 8<br />

Distribution <strong>of</strong> index <strong>an</strong>d strength properties in black <strong>clays</strong><br />

8.1 Natural moisture content<br />

The distribution <strong>of</strong> natural moisture content <strong>of</strong> soils at depths <strong>of</strong> less th<strong>an</strong> 0,50m as well as<br />

those at 0,50m <strong>an</strong>d greater, is summarised diagrammatically in Figures (8.1a & b),<br />

respectively.<br />

Natural moisture content varation; < 0,50m depth<br />

natural moisture content<br />

(%)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Figure 8.1a. Natural moisture content variation across the study area for<br />

Depths less th<strong>an</strong> 0,50m (north-south dist<strong>an</strong>ce: 4000 m).<br />

Field pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E show moisture contents <strong>of</strong> 19-34%, 20-24%, 18-26%, 18-27%<br />

<strong>an</strong>d 21-25% for depths <strong>of</strong> less th<strong>an</strong> 0,50m; <strong>an</strong>d 23-32%, 20-29%, 17-26%, 21-29% <strong>an</strong>d 21-<br />

29% for depths <strong>of</strong> 0,50m <strong>an</strong>d greater, respectively.<br />

Natural moisture content variation; 0,50m <strong>an</strong>d greater<br />

Natural moisture content<br />

(%)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Figure 8.1b. Natural moisture content variation across the study area for the<br />

depth <strong>of</strong> 0,50m <strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).


156<br />

Table 8.1. Values <strong>of</strong> index properties <strong>of</strong> soils along field pr<strong>of</strong>iles in this study.<br />

Pr<strong>of</strong>ile<br />

Depth<br />

(m)<br />

Sample No.<br />

Sta.<br />

position (m)<br />

Wn<br />

(%) LL (%) PI (%)<br />

LS<br />

(%)<br />

FS<br />

(%)<br />

Depth<br />

(m)<br />

Sample No.<br />

Wn<br />

(%) LL (%) PI (%)<br />

SA1-30cm 0 32 78 41 25 135 SA1-50cm 30 75 39 23 100<br />

SA2-<br />

A < 0,50 SA2-30cm 250 34 93 54 27 135 ≥ 0,50 70/105cm 32 89 55 26 145<br />

SA28-30cm 6750 26 85 45 25 125 SA28-50cm 27 89 47 25 135<br />

SA37-<br />

SA37-30cm 9000 19 81 43 24 115<br />

50/80cm 23 84 39 26 115<br />

SA41-30cm 10000 23 89 47 26 135 SA41-50cm 25 90 49 24 120<br />

SB1-30cm 0 23 82 47 23 125 SB1-50/70cm 29 87 50 25 140<br />

B < 0,50 SB5-30cm 1000 24 81 47 23 120 ≥ 0,50 SB5-50/70cm 29 88 50 25 140<br />

SB7-30cm 1500 23 83 48 24 130 SB7-50/70cm 28 87 49 25 138<br />

SB18-30cm 4250 20 82 43 23 120 SB18-50cm 20 72 39 21 120<br />

SB27-30cm 6500 22 85 47 24 130 SB27-50cm 22 88 49 25 130<br />

SB31-30cm 7500 23 87 49 24 130 SB31-50cm 23 86 48 25 130<br />

SB41-30cm 10000 23 89 44 26 145 SB41-50cm 24 91 47 25 145<br />

SB42-30cm 10250 22 88 44 25 145 SB42-50cm 21 91 47 26 145<br />

SC1-30cm 0 24 82 48 23 125 SC1-50cm 26 85 48 24 135<br />

C < 0,50 SC5-30cm 1000 23 83 48 23 125 ≥ 0,50 SC5-50cm 25 84 47 24 135<br />

SC9-30cm 2000 25 84 48 24 130 SC9-50cm 26 85 48 24 135<br />

SC17-30cm 4000 21 76 41 25 105 SC17-50cm 22 81 48 24 110<br />

SC25-30cm 6000 26 88 49 25 125 SC25-50cm 26 89 48 25 145<br />

SC29-30cm 7000 21 82 49 25 110 SC29-50cm 20 80 43 26 110<br />

SC33-30cm 8000 20 82 49 25 120 SC33-50cm 21 83 50 26 120<br />

SC41-30cm 10000 18 85 47 29 125 SC41-50cm 17 84 47 29 125<br />

SD1-30cm 0 27 82 46 26 125 SD1-50/80cm 29 87 48 27 135<br />

D < 0,50 SD5-30cm 1000 26 84 47 26 125 ≥ 0,50 SD5-50/70cm 29 88 49 27 135<br />

SD7-30cm 1500 27 84 48 27 110 SD7-50/70cm 29 89 51 26 138<br />

SD17-30cm 4000 20 84 44 24 120 SD17-50cm 22 84 49 25 130<br />

SD25-30cm 6000 22 80 39 25 125 SD25-50cm 21 86 45 25 135<br />

SD29-30cm 7000 18 85 47 27 125 SD29-50cm 22 86 51 26 145<br />

SD33-30cm 8000 20 85 47 27 125 SD33-50cm 21 86 50 25 145<br />

SD41-30cm 10000 23 85 46 26 130 SD41-50cm 25 90 49 26 135<br />

SE1-30cm 0 24 82 48 23 125 SE1-50/70cm 29 88 50 25 140<br />

E < 0,50 SE5-30cm 1000 25 85 48 24 125 ≥ 0,50 SE5-50/70cm 28 89 49 26 138<br />

SE13-30cm 3000 21 86 47 24 125 SE13-50cm 24 89 49 25 130<br />

SE21-30cm 5000 25 88 49 25 125 SE21-50cm 26 90 49 26 145<br />

SE29-30cm 7000 22 81 46 25 120 SE29-50cm 23 83 47 26 120<br />

SE37-30cm 9000 21 85 48 24 125 SE37-50cm 22 85 48 25 130<br />

SE41-30cm 10000 22 87 47 25 135 SE41-50cm 21 88 48 26 135<br />

max 34 93 54 29 145 max 32 91 55 29 145<br />

Statistical < 0,50 min 18 76 39 23 105 ≥ 0,50 min 17 72 39 21 100<br />

<strong>an</strong>alysis r<strong>an</strong>ge 16 17 15 6 40 r<strong>an</strong>ge 15 19 16 8 45<br />

medi<strong>an</strong> 23 84 47 25 125 medi<strong>an</strong> 25 87 48 25 135<br />

mode 23 85 47 25 125 mode 29 89 49 25 135<br />

me<strong>an</strong> 23 84 47 25 126 me<strong>an</strong> 25 86 48 25 132<br />

std 3,37 3,30 2,78 1,38 8,26 std 3,56 4,11 3,28 1,29 11,33<br />

128,2<br />

var 11,36 10,90 7,74 1,89 68,25 var 12,69 16,86 10,76 1,67 8<br />

cov 9,71 8,28 5,60 1,05 35,52 cov 9,71 8,28 5,60 1,05 35,52<br />

correlation 0,83 0,63 0,63 0,61 0,39 correlation 0,83 0,63 0,63 0,61 0,39<br />

There is a general decrease <strong>of</strong> moisture content <strong>of</strong> soils west-eastwards across the study area<br />

at depths <strong>of</strong> less th<strong>an</strong> 0,50m, as well as those at 0,50m <strong>an</strong>d more. This could be attributed to a<br />

similar variation in the total thickness <strong>of</strong> soils which also decreases west-eastwards across the<br />

study area, from depths <strong>of</strong> 1,50-2,0m in the western parts to 0,50-0,70m in the eastern parts <strong>of</strong><br />

LS<br />

(%)<br />

FS<br />

(%)


157<br />

the study area, respectively. Representation <strong>of</strong> this moisture variation in contour form is given<br />

in Fig. (8.2).<br />

Natural moisture (Wn) variation (< 0,5 m)<br />

3234 26<br />

1°19´S 4000<br />

19 23<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

27 26 27<br />

3000<br />

20 22 18 20 23<br />

23 24 23 20 22 23 23<br />

2000<br />

West<br />

24 25 21 25 22 21 22<br />

1000<br />

24 23 25 21 26<br />

1°21.5´S 0<br />

21 20 18<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

Natural moisture (Wn) variation (> 0,5 m)<br />

36°49´E<br />

3032 27 23 25<br />

1°19´S 4000<br />

Wn%<br />

32<br />

31<br />

30<br />

29<br />

28<br />

27<br />

26<br />

25<br />

24<br />

23<br />

22<br />

21<br />

20<br />

19<br />

18<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

29 29 29 22 21 22 21 25<br />

3000<br />

29 29 28 20 22 23 24<br />

2000<br />

West 29 28 24 26 23 22 21<br />

1000<br />

26 25 26 22 26 20 21 17<br />

0<br />

1°21.5´S<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

Wn%<br />

30<br />

Fig. 8.2. Contoured representation <strong>of</strong> distribution <strong>an</strong>d variation <strong>of</strong> natural moisture content<br />

in black <strong>clays</strong>.<br />

The area is dotted with patches <strong>of</strong> relatively low moisture content (Fig. 8.2) corresponding to<br />

increasingly lateritic black <strong>clays</strong>; as well as those <strong>of</strong> relatively high moisture content in<br />

swampy parts <strong>an</strong>d vicinity <strong>of</strong> stream <strong>an</strong>d/ or river valleys.


158<br />

The distribution <strong>an</strong>d variation <strong>of</strong> natural moisture content at depths <strong>of</strong> less th<strong>an</strong> 0,50m across<br />

the study area are characterised by a maximum <strong>an</strong>d minimum <strong>of</strong> 34% <strong>an</strong>d 18%, giving a r<strong>an</strong>ge<br />

<strong>of</strong> 16% (Table 8.1). Also characteristic are medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> 23% as well as<br />

a st<strong>an</strong>dard deviation <strong>of</strong> 3,37 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 11,36. The distribution is similar at depths <strong>of</strong><br />

0,50m <strong>an</strong>d greater, with maximum <strong>an</strong>d minimum values <strong>of</strong> 32% <strong>an</strong>d 17% giving a r<strong>an</strong>ge <strong>of</strong><br />

15% (Table 8.1). However, a slight increase in moisture content with depth <strong>of</strong> soils occurs<br />

(due to reduced drying effects), <strong>an</strong>d is reflected by a medi<strong>an</strong> <strong>of</strong> 25%, mode <strong>of</strong> 29% <strong>an</strong>d me<strong>an</strong><br />

<strong>of</strong> 25%. Also calculated for this depth are a st<strong>an</strong>dard deviation <strong>of</strong> 3,56 <strong>an</strong>d a vari<strong>an</strong>ce <strong>of</strong><br />

12,69.<br />

The variation <strong>of</strong> natural moisture content <strong>of</strong> soils at the two depth intervals across the study<br />

area is generally similar, so that the two sets <strong>of</strong> data give a covari<strong>an</strong>ce <strong>of</strong> 9,71 <strong>an</strong>d a strong<br />

correlation <strong>of</strong> 0,83 (Table 8.1).<br />

The upper 50cm <strong>of</strong> red soils were found to exhibit a relatively elevated moisture content <strong>of</strong><br />

38-39%, most probably a result <strong>of</strong> recent rainfall. However, depths greater th<strong>an</strong> 0,50m are<br />

generally characterised by a moisture content <strong>of</strong> 24-26%.<br />

8.2 Liquid limit<br />

The distribution <strong>of</strong> liquid limit values across the study area is presented in Figures (8.3a & b),<br />

which show a concentration about a me<strong>an</strong> <strong>of</strong> 84% <strong>an</strong>d 86%, respectively.<br />

Liquid limit variation; < 0,50m depth<br />

Liquid limit (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Figure 8.3a. Liquid limit variation across the study area for the depth <strong>of</strong> less th<strong>an</strong><br />

0,50m (north-south dist<strong>an</strong>ce: 4000 m).


159<br />

Liquid limit variation; 0,50m <strong>an</strong>d greater<br />

Liquid limit (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Figure 8.3b. Liquid limit variation across the study area for the depth <strong>of</strong><br />

0,50m <strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).<br />

The distribution is generally similar along the five field pr<strong>of</strong>iles, both at depths <strong>of</strong> less th<strong>an</strong><br />

0,50m <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d deeper. The pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E exhibit liquid limit<br />

values <strong>of</strong> 78-93%, 81-89%, 76-88%, 80-85% <strong>an</strong>d 81-88% for depths <strong>of</strong> less th<strong>an</strong> 0,50m; <strong>an</strong>d<br />

75-90%, 72-91%, 80-89%, 84-90% <strong>an</strong>d 83-90% for depths <strong>of</strong> 0,50m <strong>an</strong>d greater.<br />

The distribution <strong>an</strong>d variation <strong>of</strong> liquid limit at depths <strong>of</strong> 0,50m <strong>an</strong>d those at 0,50m <strong>an</strong>d<br />

greater across the study area are generally similar, being characterised by maximum <strong>an</strong>d<br />

minimum values <strong>of</strong> 93% <strong>an</strong>d 76% with a r<strong>an</strong>ge <strong>of</strong> 17%; <strong>an</strong>d 91% <strong>an</strong>d 72% with a r<strong>an</strong>ge <strong>of</strong><br />

19%, respectively (Table 8.1).


160<br />

Liquid limit (LL) variation (< 0,5 m)<br />

36°55´E<br />

LL (%)<br />

1°19´S<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

West<br />

1°21.5´S<br />

36°49´E<br />

LL%<br />

88<br />

87.5<br />

87<br />

86.5<br />

86<br />

85.5<br />

85<br />

84.5<br />

84<br />

83.5<br />

83<br />

82.5<br />

82<br />

81.5<br />

81<br />

80.5<br />

80<br />

79.5<br />

79<br />

78.5<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

Fig. 8.4a. Block diagram showing distribution <strong>an</strong>d variation <strong>of</strong> liquid limit values at depths <strong>of</strong><br />

less th<strong>an</strong> 0,5m in black <strong>clays</strong>.<br />

A contoured representation <strong>of</strong> the distribution is shown in Figures (8.4a & b) in which<br />

isolated patches <strong>of</strong> relatively low liquid limits correspond to zones <strong>of</strong> more lateritic <strong>clays</strong><br />

<strong>an</strong>d/or swampy environments. On the whole, however, there is a slight tendency <strong>of</strong> liquid<br />

limit values to increase with depth. This could be attributed to the relatively higher org<strong>an</strong>ic<br />

matter content <strong>of</strong> black <strong>clays</strong> for depths <strong>of</strong> less th<strong>an</strong> 0,50m, <strong>an</strong>d which has the effect <strong>of</strong><br />

inhibiting the plasticity <strong>of</strong> soils. As a result, soil depths <strong>of</strong> less th<strong>an</strong> 0,50m show medi<strong>an</strong>,<br />

mode <strong>an</strong>d me<strong>an</strong> liquid limit values <strong>of</strong> 84%, 85%, <strong>an</strong>d 84%, respectively, as well as st<strong>an</strong>dard<br />

deviation <strong>of</strong> 3,30 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 10,90; while soil depths <strong>of</strong> 0,50m <strong>an</strong>d greater, exhibit<br />

relatively larger medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> 87%, 89% <strong>an</strong>d 86%, respectively. The<br />

latter soil depths are also characterised by st<strong>an</strong>dard deviation <strong>of</strong> 4,11 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 16,86.<br />

On the whole, however, the two sets <strong>of</strong> data at the two different depth intervals are generally<br />

comparable <strong>an</strong>d in good agreement, this being indicated by a covari<strong>an</strong>ce <strong>of</strong> 8,28 <strong>an</strong>d a fairly<br />

strong correlation <strong>of</strong> 0,63 (Table 8.1).


161<br />

Liquid limit (LL) variation (> 0,5 m)<br />

36°55´E<br />

89<br />

88<br />

87<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

LL%<br />

90<br />

LL (%)<br />

South<br />

84<br />

85<br />

86<br />

83<br />

82<br />

1°19´S<br />

81<br />

80<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

West<br />

1°21.5´S<br />

36°49´E<br />

Fig. 8.4b. Block diagram showing distribution <strong>an</strong>d variation <strong>of</strong> liquid limit values at depths <strong>of</strong><br />

more th<strong>an</strong> 0,5m in black <strong>clays</strong>.<br />

76<br />

77<br />

78<br />

79<br />

Liquid limit values obtained for the red soils in this study are comparatively lower, <strong>an</strong>d r<strong>an</strong>ge<br />

from 48% to 57%.<br />

8.3 Plasticity index<br />

The variation <strong>of</strong> plasticity index <strong>of</strong> soils along the five field pr<strong>of</strong>iles <strong>of</strong> this study is<br />

represented diagrammatically in Figures (8.5a & b), for depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong><br />

0,50m <strong>an</strong>d greater, respectively.


162<br />

Plasticity index variation; < 0,50m depth<br />

Plasticity index (%)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Figure 8.5a. Plasticity index variation across the study area for the depth<br />

<strong>of</strong> less th<strong>an</strong> 0,50m (north-south dist<strong>an</strong>ce: 4000 m).<br />

Pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E show plasticity indices <strong>of</strong> 41-54%, 43-49%, 41-49%, 39-48% <strong>an</strong>d<br />

46-49% for depths <strong>of</strong> less th<strong>an</strong> 0,50m; <strong>an</strong>d 39-55%, 39-50%, 43-50%, 45-51% <strong>an</strong>d 47-50%<br />

for depths <strong>of</strong> 0,5m <strong>an</strong>d greater, respectively.<br />

Plasticity index variation; 0,50m <strong>an</strong>d greater<br />

60<br />

Plasticity index (%)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Figure 8.5b. Plasticity index variation across the study area for the depth <strong>of</strong><br />

0,50m <strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).<br />

The plasticity index <strong>of</strong> soils is similarly distributed across the study area for both depths <strong>of</strong><br />

less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d deeper, exhibiting maximum <strong>an</strong>d minimum values <strong>of</strong><br />

54% <strong>an</strong>d 39% with a r<strong>an</strong>ge <strong>of</strong> 15%; <strong>an</strong>d 55% <strong>an</strong>d 39% with a r<strong>an</strong>ge <strong>of</strong> 16%, respectively<br />

(Table 8.1). Soil depths <strong>of</strong> less th<strong>an</strong> 0,50m are also characterised by a medi<strong>an</strong>, mode <strong>an</strong>d<br />

me<strong>an</strong> value <strong>of</strong> 47%, as well as a st<strong>an</strong>dard deviation <strong>of</strong> 2,78 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 7,74. Similar <strong>an</strong>d<br />

closely comparable parameters were calculated for depths <strong>of</strong> 0,50m <strong>an</strong>d greater, in terms <strong>of</strong><br />

medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> 48%, 49% <strong>an</strong>d 48%, respectively. A st<strong>an</strong>dard deviation <strong>of</strong><br />

3,28 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 10,76 were also computed for the latter depth interval. A summary <strong>of</strong><br />

the distribution in the form <strong>of</strong> contour maps for the two depth intervals is presented in Fig.


163<br />

(8.6). The few isolated patches <strong>of</strong> relatively low plasticity values imply areas <strong>of</strong> increasingly<br />

lateritic <strong>clays</strong> <strong>an</strong>d/ or swampy environments .<br />

PI%<br />

Plasticitx index (PI) variation (< 0,5m)<br />

1°19´S 41 54 45 43 47<br />

4000<br />

(m)<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

46<br />

3000<br />

47 48 44 39 47 47 46<br />

47 47 48 43 47 49 44<br />

2000<br />

West<br />

48 48 47 49 46 48 47<br />

1000<br />

1°21.5´S48 48 48 41 49 49 49<br />

0<br />

47<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 (m)<br />

36°49´E<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

49<br />

48.5<br />

48<br />

47.5<br />

47<br />

46.5<br />

%<br />

46<br />

45.5<br />

45<br />

44.5<br />

44<br />

43.5<br />

43<br />

42.5<br />

42<br />

41.5<br />

41<br />

40.5<br />

Plasticity index (PI) variation (> 0,5m)<br />

1°19´S 39 55 47 39 49<br />

4000<br />

(m)<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

48 49 51 49 45 51 50 49<br />

3000<br />

50 50 49 39 49<br />

2000<br />

48 47<br />

West<br />

50 49 49 49 47 48 48<br />

1000<br />

1°21.5´S48 47 48 48 48 43 50<br />

0<br />

47<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 (m)<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

41.5<br />

42<br />

42.5<br />

43<br />

43.5<br />

44<br />

44.5<br />

45<br />

45.5<br />

46<br />

46.5<br />

47<br />

47.5<br />

48<br />

48.5<br />

49<br />

49.5<br />

50<br />

PI%<br />

50.5<br />

Figure 8.6. A contoured representation <strong>of</strong> the distribution <strong>an</strong>d variation <strong>of</strong> plasticity index<br />

values in black <strong>clays</strong>.<br />

The two depth intervals show more or less similar plasticity values <strong>an</strong>d parameters, implying<br />

a more or less uniform <strong>an</strong>d/ or homogeneous character <strong>of</strong> the black <strong>clays</strong>, both across the<br />

study area <strong>an</strong>d with depth. This is further evidenced by a covari<strong>an</strong>ce <strong>of</strong> 5,60 <strong>an</strong>d a fairly<br />

strong correlation <strong>of</strong> 0,63 between the two sets <strong>of</strong> data obtained for the two depth intervals<br />

(Table 8.1).<br />

The plasticity indices obtained for the red soils in this study are relatively low, r<strong>an</strong>ging from<br />

18-22%.


164<br />

8.4 Linear shrinkage<br />

The spatial variation <strong>an</strong>d distribution <strong>of</strong> values <strong>of</strong> linear shrinkage determined in this study are<br />

presented in Table (8.1) <strong>an</strong>d Figures (8.7a & b).<br />

Linear shrinkage variation; < 0,5m depth<br />

30<br />

Linear shrinkage (%)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Fig. 8.7a. Linear shrinkage variation across the study area for the depth <strong>of</strong> less th<strong>an</strong> 0,50m (north-south dist<strong>an</strong>ce: 4000 m).<br />

Black <strong>clays</strong> <strong>of</strong> less th<strong>an</strong> 0,50m depth generally exhibit little variation in linear shrinkage<br />

values along each <strong>of</strong> the five field pr<strong>of</strong>iles across the study area (Fig. 8.7a); with percentage<br />

values <strong>of</strong> 24-27, 23-26, 23-29, 24-27 <strong>an</strong>d 23-25 determined for pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E,<br />

respectively. On the whole, soil thicknesses in this depth interval across the study area exhibit<br />

linear shrinkage values <strong>of</strong> between 23-29%, with a r<strong>an</strong>ge <strong>of</strong> 6% as well as mode, medi<strong>an</strong> <strong>an</strong>d<br />

me<strong>an</strong> values <strong>of</strong> 25%. The distribution is also characterised by a st<strong>an</strong>dard deviation <strong>of</strong> 1,38<br />

<strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 1,89.<br />

Linear shrinkage variation; 0,50m <strong>an</strong>d more<br />

30<br />

25<br />

Linear shrinkage (%)<br />

20<br />

15<br />

10<br />

5<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Figure 8.7b. Linear shrinkage variation across the study area at depths <strong>of</strong> 0,5m <strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).


165<br />

Similarly, soil thicknesses <strong>of</strong> black <strong>clays</strong> with depths equal to <strong>an</strong>d/ or greater th<strong>an</strong> 0,50m show<br />

linear shrinkage values generally concentrated about 25% (Figure 8.7b); with percentage<br />

values <strong>of</strong> 23-26, 21-26, 24-29, 25-27 <strong>an</strong>d 25-26 recorded for pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E,<br />

respectively. On the whole across the study area, values <strong>of</strong> 21-29% exhibiting a r<strong>an</strong>ge <strong>of</strong> 8%<br />

as well as a mode, medi<strong>an</strong> <strong>an</strong>d me<strong>an</strong> <strong>of</strong> 25% have been recorded for the depth. A st<strong>an</strong>dard<br />

deviation <strong>of</strong> 1,29 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 1,67, have also been calculated. Fig. (8.8) shows a<br />

contoured distribution pattern <strong>of</strong> linear shrinkage at the two depth intervals.<br />

Linear shrinkage (LS) variation (< 0,5 m)<br />

2527 25 24 26<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

26 26 27 24 25 27 27 26<br />

3000<br />

23 23 24 23 24 24 26<br />

2000<br />

West<br />

23 24 24 25 25 24 25<br />

1000<br />

23 23 24 25 25 25 25 29<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

LS (%)<br />

27.5<br />

27<br />

26.5<br />

26<br />

25.5<br />

25<br />

24.5<br />

24<br />

23.5<br />

23<br />

Linear shrinkage (LS) variation (> 0,5 m)<br />

2326 25 26 24<br />

1°19´S 4000<br />

22.5<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

27 27 26 25 25 26 25 26<br />

3000<br />

25 25 25 21 25 25 25<br />

2000<br />

West 25 26 25 26 26 25 26<br />

1000<br />

1°21.5´S<br />

24 24 24 24 25 26 26 29<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

27<br />

26.5<br />

26<br />

25.5<br />

25<br />

24.5<br />

24<br />

23.5<br />

23<br />

22.5<br />

22<br />

Figure 8.8. A contoured distribution pattern <strong>of</strong> linear shrinkage in black <strong>clays</strong>.<br />

27.5<br />

LS (%)<br />

A comparison <strong>of</strong> linear shrinkage values <strong>of</strong> black <strong>clays</strong> for depths <strong>of</strong> less th<strong>an</strong> 0,50m with<br />

those obtained for depths equal to/ <strong>an</strong>d greater th<strong>an</strong> 0,50m yield a moderately strong<br />

correlation <strong>of</strong> 0,61 <strong>an</strong>d a covari<strong>an</strong>ce <strong>of</strong> 1,05. In addition, there is a strong agreement between<br />

values <strong>of</strong> linear shrinkage obtained for the two depth intervals in terms <strong>of</strong> comparable<br />

respective parameters <strong>of</strong> mode, medi<strong>an</strong>, me<strong>an</strong>, st<strong>an</strong>dard deviation <strong>an</strong>d vari<strong>an</strong>ce. This serves to


166<br />

strongly suggest that the whole soil thickness <strong>of</strong> black <strong>clays</strong> <strong>of</strong> the current study area is more<br />

or less homogenous <strong>an</strong>d uniform in terms <strong>of</strong> linear shrinkage <strong>an</strong>d, possibly, other index <strong>an</strong>d/or<br />

<strong>engineering</strong> properties.<br />

Linear shrinkage values obtained for the red soils are comparatively low <strong>an</strong>d r<strong>an</strong>ge from 10-<br />

11%.<br />

8.5 Free swell<br />

The distribution <strong>of</strong> free swell values along the five field pr<strong>of</strong>iles in this study is presented in<br />

Figure (8.9). Pr<strong>of</strong>iles A, B, C, D <strong>an</strong>d E show free swell values <strong>of</strong> 115-135%, 120-145%, 105-<br />

130%, 110-130 <strong>an</strong>d 120-135% for depths <strong>of</strong> less th<strong>an</strong> 0,50m concentrated about a me<strong>an</strong> value<br />

<strong>of</strong> 126%; <strong>an</strong>d 100-145%, 120-145%, 110-145%, 130-145% <strong>an</strong>d 120-145% for depths 0f 0,5m<br />

<strong>an</strong>d over, concentrated about a me<strong>an</strong> value <strong>of</strong> 132%, respectively. The complete set <strong>of</strong> data is<br />

given in Table (8.1).<br />

Distribution <strong>of</strong> free swell values at a depth <strong>of</strong> less th<strong>an</strong> 0,50m across the study area is<br />

characterised by minimum <strong>an</strong>d maximun values <strong>of</strong> 105% <strong>an</strong>d 145% respectively, giving a<br />

r<strong>an</strong>ge <strong>of</strong> 40%, medi<strong>an</strong> <strong>an</strong>d mode <strong>of</strong> 125%, me<strong>an</strong> <strong>of</strong> 126% as well as a st<strong>an</strong>dard deviation <strong>of</strong><br />

8,26 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 68,25.<br />

Free swell variation; < 0,50m depth<br />

Free swell (%)<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Figure 8.9a. Free swell variation across the study area at depths <strong>of</strong> less th<strong>an</strong><br />

0,5m (north-south dist<strong>an</strong>ce: 4000 m).<br />

On the other h<strong>an</strong>d, black <strong>clays</strong> at depths <strong>of</strong> 0,50m <strong>an</strong>d over exhibit free swell values <strong>of</strong><br />

between 100% <strong>an</strong>d 145% with a r<strong>an</strong>ge <strong>of</strong> 45%, across the study area. A medi<strong>an</strong> <strong>an</strong>d mode <strong>of</strong><br />

135%, me<strong>an</strong> <strong>of</strong> 132%, st<strong>an</strong>dard deviation <strong>of</strong> 11,33 <strong>an</strong>d vari<strong>an</strong>ce <strong>of</strong> 128,28 were also<br />

calculated for the depth.


167<br />

Free swell variation; 0,50m depth <strong>an</strong>d greater<br />

Free swell (%)<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Figure 8.9b. Free swell variation across the study area at depths <strong>of</strong> 0,50m<br />

<strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).<br />

The two sets <strong>of</strong> data obtained for depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d over, are<br />

generally comparable <strong>an</strong>d similar in their distribution in terms <strong>of</strong> maximum <strong>an</strong>d minimum<br />

values as well as their r<strong>an</strong>ge. A slight variation is, however, exhibited in terms <strong>of</strong> the medi<strong>an</strong>,<br />

mode <strong>an</strong>d me<strong>an</strong> values as well as the st<strong>an</strong>dard deviation <strong>an</strong>d vari<strong>an</strong>ce. As a result, a<br />

covari<strong>an</strong>ce <strong>of</strong> 35,52 <strong>an</strong>d relatively weak correlation <strong>of</strong> 0,39 has also been recorded for the two<br />

sets <strong>of</strong> data. Relatively lower values <strong>of</strong> free swell <strong>an</strong>d/ or swelling capability in upper soil<br />

depths could be attributed to increased org<strong>an</strong>ic matter content.<br />

Fig. (8.10) shows a contoured representation <strong>of</strong> the above described distribution, in which<br />

patches <strong>of</strong> low swelling capabilities are associated with lateritic <strong>clays</strong>; as well as <strong>clays</strong> <strong>of</strong><br />

swampy environments <strong>an</strong>d/ or those adjacent to river valleys <strong>an</strong>d stream ch<strong>an</strong>nels with<br />

impeded drainage <strong>an</strong>d relatively higher org<strong>an</strong>ic matter content.<br />

Free swell values obtained for red soils in this study are comparatively low <strong>an</strong>d r<strong>an</strong>ge from<br />

15-20%.


168<br />

Free swell (FS) variation (< 0,5 m)<br />

1°19´S 135135 125 115 135<br />

4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

125 125 110 120 125 125 125 130<br />

3000<br />

125 120 130 120 130 130 145<br />

2000<br />

West 125 125 125 125 120 125 135<br />

1000<br />

125 125 130 105 125 110 120 125<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

Free swell (FS) variation (> 0,5 m)<br />

1°19´S 100 145 135 115 120<br />

4000<br />

FS%<br />

140<br />

138<br />

136<br />

134<br />

132<br />

130<br />

128<br />

126<br />

124<br />

122<br />

120<br />

118<br />

116<br />

114<br />

112<br />

110<br />

108<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

135 135 138 130 135 145 145 135<br />

3000<br />

140 140 138 120 130 130 145<br />

2000<br />

West 140 138 130 145 120 130 135<br />

1000<br />

135 135 135 110<br />

145<br />

110 120 125<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

114<br />

116<br />

118<br />

120<br />

122<br />

124<br />

126<br />

128<br />

Figure 8.10. Contour maps showing distribution <strong>of</strong> free swell values <strong>an</strong>d variation <strong>of</strong> swelling<br />

capability in black <strong>clays</strong>.<br />

130<br />

132<br />

134<br />

136<br />

138<br />

FS%<br />

140<br />

8.6 Clay <strong>an</strong>d fine fraction<br />

The distribution <strong>an</strong>d variation <strong>of</strong> percentage clay <strong>an</strong>d fines (clay + silt) content <strong>of</strong> black <strong>clays</strong><br />

across the study area is summarised in Table (8.2) in the order <strong>of</strong> pr<strong>of</strong>iles. The variation <strong>of</strong><br />

clay fraction at depths <strong>of</strong> less th<strong>an</strong> 0,50m is characterised by maximum <strong>an</strong>d minimum values<br />

<strong>of</strong> 37% <strong>an</strong>d 13% giving a r<strong>an</strong>ge <strong>of</strong> 24%; as well as medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values 29%, 31%<br />

<strong>an</strong>d 27%, respectively. A similar variation <strong>of</strong> clay fraction occurs at depths <strong>of</strong> 0,50m <strong>an</strong>d<br />

greater, where closely comparable maximum <strong>an</strong>d minimum values <strong>of</strong> 44% <strong>an</strong>d 14% giving a<br />

r<strong>an</strong>ge <strong>of</strong> 30%; as well as medi<strong>an</strong>/ mode <strong>of</strong> 27% <strong>an</strong>d me<strong>an</strong> <strong>of</strong> about 26% have been registered.<br />

Closely agreeable st<strong>an</strong>dard deviation values <strong>of</strong> 6,09 <strong>an</strong>d 6,54; as well a strong correlation <strong>of</strong><br />

0,82 for the two sets <strong>of</strong> data serve as further evidence for similarity <strong>of</strong> variation <strong>of</strong> clay


169<br />

fraction at the two depth intervals across the study area. The distribution <strong>an</strong>d variation <strong>of</strong><br />

percent clay in black <strong>clays</strong> across the study area is also summarised in contour form in Fig.<br />

(8.11).<br />

Table 8.2. Values <strong>of</strong> clay <strong>an</strong>d fines fractions <strong>of</strong> soils along field pr<strong>of</strong>iles in this study.<br />

Pr<strong>of</strong>ile Depth (m) Sample No. Sta. position (m) PI (%) % clay % fines Depth (m) Sample No. PI (%) % clay % fines<br />

SA1-30cm 0 41 18 90 SA1-50cm 39 16 72<br />

A < 0,50 SA2-30cm 250 54 26 90 ≥ 0,50 SA2-70/105cm 55 25 87<br />

SA28-30cm 6750 45 23 94 SA28-50cm 47 18 88<br />

SA37-30cm 9000 43 32 89 SA37-50/80cm 39 33 91<br />

SA41-30cm 10000 47 37 96 SA41-50cm 49 44 89<br />

SB1-30cm 0 47 30 90 SB1-50/70cm 50 26 91<br />

B < 0,50 SB5-30cm 1000 47 29 89 ≥ 0,50 SB5-50/70cm 50 26 91<br />

SB7-30cm 1500 48 31 89 SB7-50/70cm 49 27 91<br />

SB18-30cm 4250 43 13 84 SB18-50cm 39 14 88<br />

SB27-30cm 6500 47 32 94 SB27-50cm 49 33 92<br />

SB31-30cm 7500 49 35 90 SB31-50cm 48 30 93<br />

SB41-30cm 10000 44 20 88 SB41-50cm 47 22 86<br />

SB42-30cm 10250 44 23 88 SB42-50cm 47 21 86<br />

SC1-30cm 0 48 30 87 SC1-50cm 48 27 90<br />

C < 0,50 SC5-30cm 1000 48 31 88 ≥ 0,50 SC5-50cm 47 27 91<br />

SC9-30cm 2000 48 31 89 SC9-50cm 48 27 89<br />

SC17-30cm 4000 41 29 88 SC17-50cm 48 27 86<br />

SC25-30cm 6000 49 23 94 SC25-50cm 48 18 89<br />

SC29-30cm 7000 49 28 91 SC29-50cm 43 26 92<br />

SC33-30cm 8000 49 27 91 SC33-50cm 50 28 91<br />

SC41-30cm 10000 47 16 74 SC41-50cm 47 19 85<br />

SD1-30cm 0 46 32 91 SD1-50/80cm 48 27 90<br />

D < 0,50 SD5-30cm 1000 47 31 91 ≥ 0,50 SD5-50/70cm 49 27 90<br />

SD7-30cm 1500 48 32 90 SD7-50/70cm 51 26 90<br />

SD17-30cm 4000 44 13 84 SD17-50cm 49 21 84<br />

SD25-30cm 6000 39 24 84 SD25-50cm 45 14 74<br />

SD29-30cm 7000 47 27 90 SD29-50cm 51 28 92<br />

SD33-30cm 8000 47 29 90 SD33-50cm 50 28 91<br />

SD41-30cm 10000 46 35 92 SD41-50cm 49 41 89<br />

SE1-30cm 0 48 31 89 SE1-50/70cm 50 27 91<br />

E < 0,50 SE5-30cm 1000 48 31 89 ≥ 0,50 SE5-50/70cm 49 27 91<br />

SE13-30cm 3000 47 33 93 SE13-50cm 49 34 92<br />

SE21-30cm 5000 49 24 94 SE21-50cm 49 20 88<br />

SE29-30cm 7000 46 27 92 SE29-50cm 47 29 92<br />

SE37-30cm 9000 48 33 90 SE37-50cm 48 31 92<br />

SE41-30cm 10000 47 19 89 SE41-50cm 48 19 85<br />

max 54 37 96 max 55 44 93<br />

Statistical < 0,50 min 39 13 74 ≥ 0,50 min 39 14 72<br />

<strong>an</strong>alysis r<strong>an</strong>ge 15 24 22 r<strong>an</strong>ge 16 30 21<br />

medi<strong>an</strong> 47 29 90 medi<strong>an</strong> 48 27 90<br />

mode 47 31 90 mode 49 27 91<br />

me<strong>an</strong> 47 27 89 me<strong>an</strong> 48 26 89<br />

std 2,78 6,09 3,81 std 3,28 6,54 4,50<br />

var 7,74 37,04 14,48 var 10,76 42,82 20,25<br />

cov 31,86 5,92 cov 31,86 5,92<br />

correlation 0,82 0,36 correlation 0,82 0,36


170<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

3000<br />

2000<br />

West<br />

1000<br />

1°21.5´S<br />

1°19´S 4000<br />

3000<br />

2000<br />

West<br />

1000<br />

% clay variation (< 0,5 m)<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

% clay variation (> 0,5 m)<br />

%clay<br />

35<br />

34<br />

33<br />

32<br />

31<br />

30<br />

29<br />

28<br />

27<br />

26<br />

25<br />

24<br />

23<br />

22<br />

21<br />

20<br />

19<br />

18<br />

17<br />

16<br />

15<br />

14<br />

1°21.5´S<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

38<br />

36<br />

34<br />

32<br />

30<br />

28<br />

26<br />

24<br />

22<br />

20<br />

18<br />

16<br />

14<br />

Figure 8.11. Contour maps showing distribution <strong>an</strong>d variation <strong>of</strong> %clay in black <strong>clays</strong>.<br />

40<br />

% clay<br />

The variation <strong>of</strong> percentage fines is also identical at the two depth intervals, in terms <strong>of</strong><br />

maximum <strong>an</strong>d minimum values <strong>of</strong> 96% <strong>an</strong>d 74% (r<strong>an</strong>ge: 22%); <strong>an</strong>d 93% <strong>an</strong>d 72% (r<strong>an</strong>ge:<br />

21%), respectively; as well as medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> about 90%. St<strong>an</strong>dard<br />

deviations registered for the data sets at the two depth intervals are also <strong>of</strong> the same order <strong>of</strong><br />

magnitude, i.e. 3,81 <strong>an</strong>d 4,50, respectively. The variation along field pr<strong>of</strong>iles at the two depth<br />

intervals is also illustrated in Figure (8.12) <strong>an</strong>d a contoured representation in Fig.(8.13).


171<br />

Variation <strong>of</strong> fines (%), < 0,50m depth<br />

% fines<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Fig. 8.12a. Variation <strong>of</strong> fines fraction across the study area at depths <strong>of</strong> less th<strong>an</strong> 0,5m (north-south dist<strong>an</strong>ce: 4000 m).<br />

Variation <strong>of</strong> fines (%), 0,50m depth <strong>an</strong>d greater<br />

% fines<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

Dist<strong>an</strong>ce along pr<strong>of</strong>ile (m)<br />

Pr<strong>of</strong>ile A<br />

Pr<strong>of</strong>ile B<br />

Pr<strong>of</strong>ile C<br />

Pr<strong>of</strong>ile D<br />

Pr<strong>of</strong>ile E<br />

Fig. 8.12b. Variation <strong>of</strong> fines fraction across the study area at depths <strong>of</strong> 0,50m <strong>an</strong>d greater (north-south dist<strong>an</strong>ce: 4000 m).


172<br />

% fines variation (< 0,5 m)<br />

9090 94 89 96<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

91 91 90 84 84 90 90 92<br />

3000<br />

90 89 89 84 94 90 88<br />

2000<br />

West 89 89 93 94 92 90 89<br />

1000<br />

87 88 89 88 94 91 91<br />

74<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

%fines<br />

94<br />

93<br />

92<br />

91<br />

90<br />

89<br />

88<br />

87<br />

86<br />

85<br />

84<br />

83<br />

82<br />

81<br />

80<br />

79<br />

78<br />

77<br />

1°19´S<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

% fines variation (> 0,5 m)<br />

7287<br />

88 91 89<br />

4000<br />

90 90 90 84 74<br />

3000<br />

92 91 89<br />

91 91 91 88 92 93 86<br />

2000<br />

West 91 91 92 88 92 92 85<br />

1000<br />

1°21.5´S<br />

90 91 89 86 89 92 91 85<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

78<br />

79<br />

80<br />

81<br />

82<br />

83<br />

84<br />

85<br />

86<br />

87<br />

88<br />

89<br />

90<br />

91<br />

%fines<br />

92<br />

Figure 8.13. A contoured representation <strong>of</strong> the distribution <strong>an</strong>d variation <strong>of</strong> fines fraction in<br />

black <strong>clays</strong>.<br />

Once again, areas <strong>of</strong> relatively low percent clay <strong>an</strong>d fines fractions are associated with zones<br />

<strong>of</strong> increasingly lateritic <strong>clays</strong> (Figs. 8.11 & 8.13). Patches <strong>of</strong> lateritic <strong>clays</strong> found are therefore<br />

characterised by relatively higher contents <strong>of</strong> the coarse fraction (Fig. 8.14).


173<br />

%coarse<br />

% coarse fraction variation (< 0,5 m)<br />

1010 6 11 4<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

9 9 10 16 16 10 10 8<br />

3000<br />

10 11 11 16 6 10 12<br />

2000<br />

West 11 11 7 6 8 10 11<br />

1000<br />

13 12 11 12 6 9 9<br />

26<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

22<br />

21<br />

20<br />

19<br />

18<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

% coarse fraction variation (> 0,5 m)<br />

28 13 12 9 11<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

10 10 10 16 26<br />

8 9 11<br />

3000<br />

9 9 9 12 8 7 14<br />

2000<br />

West<br />

9 9 8 12 8 8 15<br />

1000<br />

1°21.5´S<br />

10 9 11 14 11 8 9 15<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

20<br />

19<br />

18<br />

17<br />

16<br />

15<br />

14<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

%coarse<br />

21<br />

Fig. 8.14. Distribution <strong>an</strong>d variation <strong>of</strong> coarse fraction in black <strong>clays</strong>.<br />

The red soils exhibit average clay <strong>an</strong>d fines contents <strong>of</strong> 21% <strong>an</strong>d 93%, respectively. Despite<br />

the similar proportions <strong>of</strong> fines in black <strong>clays</strong> <strong>an</strong>d red soils, the former are found to exhibit<br />

higher plasticity <strong>an</strong>d potential exp<strong>an</strong>siveness relative to the latter. The difference could be<br />

attributed to the predominating type <strong>of</strong> clay mineral composition, the high content <strong>of</strong><br />

smectites (over 90%) rendering the black <strong>clays</strong> higher swelling capabilities on wetting. The<br />

red soils are mainly kaolinite (over 80%), a clay mineral characterised by limited swelling<br />

capabilities on wetting.<br />

8.7 Shear strength<br />

Variation <strong>of</strong> strength characteristics in black <strong>clays</strong> is presented in Figures (8.15a & b) in<br />

which distribution <strong>of</strong> shear strength parameters <strong>of</strong> apparent cohesion (c´ ) <strong>an</strong>d <strong>an</strong>gle <strong>of</strong><br />

frictional resist<strong>an</strong>ce (phi or φ´ ) are given in contour form.


174<br />

1°19´S 4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

Cohesion (c`) variation (< 0,5 m)<br />

23.1 29.8<br />

47.9 44.0 45.4 32.0<br />

3000<br />

46.7 47.4 44.9 30.2<br />

2000<br />

West 46.4 42.6 25.8 26.0<br />

1000<br />

c` (kN/m²)<br />

46<br />

44<br />

42<br />

40<br />

38<br />

36<br />

34<br />

32<br />

30<br />

1°21.5´S<br />

47.2<br />

11.5 25.1<br />

28<br />

0<br />

26<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E 24<br />

22<br />

20<br />

18<br />

16<br />

1°19´S 4000<br />

Cohesion (c`) variation (> 0,5 m)<br />

35.6 41.0<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36.1 35.6 36.1 28.0<br />

3000<br />

28.4<br />

2000<br />

26.3 36.9 28.6<br />

West 40.8 35.1 36.2 29.0<br />

1000<br />

1°21.5´S<br />

31.8 33.7 34.9<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

38.5<br />

38<br />

37.5<br />

37<br />

36.5<br />

36<br />

35.5<br />

35<br />

34.5<br />

34<br />

33.5<br />

33<br />

32.5<br />

32<br />

31.5<br />

31<br />

30.5<br />

30<br />

29.5<br />

29<br />

Fig. 8.15a. Distribution <strong>an</strong>d variation <strong>of</strong> cohesion (c´ ) in black <strong>clays</strong>.<br />

c` (kN/m²)<br />

There is a slight increase in values <strong>of</strong> cohesion with increased depths (Fig. 8.15a), <strong>an</strong>d this<br />

could be attributed to a corresponding decrease <strong>of</strong> org<strong>an</strong>ic matter content with depth <strong>of</strong> soils.<br />

In practice, org<strong>an</strong>ic matter would have the effect <strong>of</strong> counteracting cohesive effects in clay<br />

soils. Similarly, the general decrease in the thickness <strong>of</strong> black <strong>clays</strong> eastwards across the study<br />

area would also me<strong>an</strong> progressively increased counteractive effects <strong>of</strong> org<strong>an</strong>ic matter content<br />

on cohesion <strong>of</strong> the <strong>clays</strong>. As a result, values <strong>of</strong> cohesion are also observed to generally<br />

decrease eastwards at <strong>an</strong>y chosen depth interval (Fig. 8.15a).<br />

The <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce is usually a more reliable parameter in characterising strength<br />

characteristics <strong>of</strong> soils (Head, 1988).


175<br />

1°19´S 22<br />

Shear <strong>an</strong>gle variation (< 0,5 m)<br />

29<br />

4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

17 16 16<br />

30<br />

3000<br />

17 17 17 30<br />

2000<br />

West 16 16 26 28<br />

1000<br />

18 22 25<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

36°55´E<br />

phi (°)<br />

29<br />

28<br />

27<br />

26<br />

25<br />

24<br />

23<br />

22<br />

21<br />

20<br />

19<br />

18<br />

17<br />

16<br />

1°19´S<br />

4000<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m)<br />

Shear <strong>an</strong>gle variation (> 0,5 m)<br />

19 19<br />

15 13 14 21<br />

3000<br />

15 16 14 22<br />

2000<br />

West<br />

11 12<br />

1000<br />

23 22<br />

14 14 22<br />

1°21.5´S 0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

36°49´E<br />

South<br />

Pr<strong>of</strong>ile dist<strong>an</strong>ce (m) 36°55´E<br />

22.5<br />

22<br />

21.5<br />

21<br />

20.5<br />

20<br />

19.5<br />

19<br />

18.5<br />

18<br />

17.5<br />

17<br />

16.5<br />

16<br />

15.5<br />

15<br />

14.5<br />

14<br />

13.5<br />

13<br />

12.5<br />

Fig. 8.15b. Distribution <strong>an</strong>d variation <strong>of</strong> <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (φ´ ) in black <strong>clays</strong>.<br />

phi (°)


176<br />

Chapter 9 Implications <strong>of</strong> index/ <strong>engineering</strong> properties <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils<br />

on construction practice<br />

9.1 Atterberg limits <strong>an</strong>d other index parameters<br />

The spatial distribution <strong>of</strong> values <strong>of</strong> Atterberg limits for black <strong>clays</strong> across the study area at<br />

soil depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d above 0,50m is presented in the last chapter (Figs. 8.3-<br />

8.6). The distribution also serves to be indicative <strong>of</strong> the possible variation <strong>of</strong> other import<strong>an</strong>t<br />

<strong>engineering</strong> soil properties across the study area. This is evidenced by plotting on the same<br />

axes results <strong>of</strong> Atterberg limits against those <strong>of</strong> free swell <strong>an</strong>d linear shrinkage (see next<br />

chapter), all <strong>of</strong> which demonstrate generally similar variation trends in their spatial<br />

distribution across the area. This points to the possibility that the Atterberg limits <strong>an</strong>d other<br />

index properties are interrelated, so that parameters <strong>of</strong> one property could be readily estimated<br />

from measured values <strong>of</strong> the other, <strong>an</strong>d vice versa. In this study, a correlation <strong>of</strong> results <strong>of</strong><br />

Atterberg limits with those <strong>of</strong> other index properties has been attempted, <strong>an</strong>d the forms <strong>an</strong>d<br />

strengths <strong>of</strong> interrelationship determined <strong>an</strong>d discussed in the next chapter.<br />

The results <strong>of</strong> Atterberg limits could also be usefully applied in the selection <strong>of</strong> soils for use<br />

as compacted fill in various types <strong>of</strong> earthworks construction. This is because <strong>clays</strong> <strong>of</strong> high<br />

plasticity are also usually associated with properties <strong>of</strong> relatively low permeability <strong>an</strong>d a<br />

tendency to consolidate over longer periods <strong>of</strong> time under load th<strong>an</strong> <strong>clays</strong> <strong>of</strong> low plasticity<br />

(Head, 1984). As a result, the high plasticity black <strong>clays</strong> involved in this study would be more<br />

difficult to compact, <strong>an</strong>d are therefore unsuitable for use as fill material; while the relatively<br />

low plasticity red soils could be preferred for the purpose.<br />

The liquid <strong>an</strong>d plastic limits also serve to indicate <strong>an</strong>d describe the consistency state <strong>of</strong> clay<br />

soil involved, i.e. the condition <strong>of</strong> the clay soils in their natural state would be dependent upon<br />

<strong>an</strong>d described by their natural moisture content in relation to these limits, as expressed by the<br />

liquidity index, LI, <strong>an</strong>d relative consistency, Cr, (Table 7.2). The black <strong>clays</strong> could be<br />

described as having a generally firm consistency; while the red soils have firm to stiff<br />

consistency. (However, swampy areas <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d forested areas <strong>of</strong> red soils were<br />

found to show generally s<strong>of</strong>t upper thicknesses <strong>of</strong> soils). As a result, red soils are expected to<br />

be comparatively more stable with respect to supporting light <strong>engineering</strong> structures th<strong>an</strong><br />

black <strong>clays</strong>. This is because the natural moisture content <strong>of</strong> soils as related to the plastic <strong>an</strong>d<br />

liquid limits through the liquidity index (LI) <strong>an</strong>d relative consistency (Cr) is also usually<br />

related to, <strong>an</strong>d serves to be indicative <strong>of</strong> the shear strength <strong>an</strong>d compressibility <strong>of</strong> soils in situ<br />

(Nelson <strong>an</strong>d Miller, 1992). A correlation <strong>of</strong> LI <strong>an</strong>d Cr values with shear strength <strong>an</strong>d<br />

consolidation parameters <strong>of</strong> soils involved in this study is presented <strong>an</strong>d discussed in<br />

respective sections <strong>of</strong> laboratory shear strength <strong>an</strong>d consolidation tests.<br />

Black <strong>clays</strong> have been found to harbour potentially high swelling (over 100% free swell) <strong>an</strong>d<br />

shrinkage (21-29% linear shrinkage) capabilities th<strong>an</strong> red soils (less th<strong>an</strong> 50% free swell;<br />

about 10% linear shrinkage). As a result, volumetric ch<strong>an</strong>ges <strong>an</strong>d/ or exp<strong>an</strong>sive movements in<br />

soils that usually accomp<strong>an</strong>y alternating seasons <strong>of</strong> wet <strong>an</strong>d dry months are likely to have a<br />

comparatively more destabilising effect on light <strong>engineering</strong> structures constructed on black<br />

<strong>clays</strong> th<strong>an</strong> those on red soils.


177<br />

9.2 Grain size distribution<br />

The results <strong>of</strong> particle size <strong>an</strong>alysis obtained in this study have been used to complement those<br />

<strong>of</strong> Atterberg limits to give a complete description, <strong>an</strong>d thereby reflect the probable physical<br />

<strong>an</strong>d/ or <strong>engineering</strong> behaviour <strong>of</strong> the clay soils. This is illustrated in Tables (7.1) <strong>an</strong>d (7.4), as<br />

well as Figure (7.2) where the colloidal activity <strong>an</strong>d probable exp<strong>an</strong>sive behaviour <strong>of</strong> the soils<br />

have been calculated <strong>an</strong>d/ or inferred from measured values <strong>of</strong> Atterberg limits (plasticity<br />

index), free swell <strong>an</strong>d percentage clay fraction. The black <strong>clays</strong> <strong>an</strong>d red soils both show high<br />

contents <strong>of</strong> the fine fraction (silt <strong>an</strong>d clay sizes) <strong>of</strong> over 70%. The higher plasticity <strong>an</strong>d<br />

potentially exp<strong>an</strong>sive behaviour exhibited by the black <strong>clays</strong> could be accounted for in terms<br />

<strong>of</strong> clay mineral composition, i.e. a higher content <strong>of</strong> smectites (over 90%) which usually<br />

exhibit high swelling capabilities on wetting. On the contrary, the red soils are predominated<br />

by the rather less exp<strong>an</strong>sive kaolinite (over 80%) which usually exhibits limited activity <strong>an</strong>d<br />

swelling capabilities on wetting.<br />

In addition, the results could be used in conjuction with those <strong>of</strong> Atterberg limits <strong>an</strong>d free<br />

swell (swelling capability) when considering the soils for use as fill material in various civil<br />

<strong>an</strong>d construction works. This is because specific civil <strong>engineering</strong> works such as<br />

emb<strong>an</strong>kments, earth dams <strong>an</strong>d/ or sub-base materials for roads <strong>an</strong>d airfield runways; all<br />

require the soils to meet certain grading specifications to provide a mech<strong>an</strong>ically stable<br />

foundation (Nelson <strong>an</strong>d Miller, 1992). The results <strong>of</strong> particle size <strong>an</strong>alysis would also serve to<br />

assess <strong>an</strong>d indicate the probable feasibility <strong>an</strong>d/ or effectiveness <strong>of</strong> dynamic compaction<br />

technique when used to improve poor ground conditions in projected sites. In the present<br />

study for inst<strong>an</strong>ce, the black <strong>clays</strong> <strong>an</strong>d red soils have been classified into soils <strong>of</strong> high <strong>an</strong>d low<br />

activity <strong>an</strong>d/ or swelling capability/ exp<strong>an</strong>siveness, respectively; on the basis <strong>of</strong> grain sizes,<br />

Atterberg limits <strong>an</strong>d free swell (Tables 7.1 <strong>an</strong>d 7.4). As a result, the black <strong>clays</strong> with probable<br />

high swelling capability/ exp<strong>an</strong>sive behaviour would generally be difficult to compact <strong>an</strong>d<br />

thus unsuitable for use as fill material; while the relatively low swelling capability/ exp<strong>an</strong>sive<br />

red soils would be preferred.<br />

Results <strong>of</strong> particle size <strong>an</strong>alysis, <strong>an</strong>d especially the clay fraction <strong>an</strong>d/ or fine fraction, have<br />

been correlated with other soil parameters <strong>an</strong>d index properties (see next chapter). This had<br />

the purpose <strong>of</strong> exploring <strong>an</strong>d establishing possible forms <strong>of</strong> interrelationships that could aid<br />

in approximately estimating <strong>an</strong>d/ or inferring for import<strong>an</strong>t soils` physical/ mech<strong>an</strong>ical <strong>an</strong>d/ or<br />

<strong>engineering</strong> behaviour relev<strong>an</strong>t to projection/ pl<strong>an</strong>ning <strong>an</strong>d construction <strong>of</strong> light structures.<br />

The results <strong>of</strong> the <strong>an</strong>alysis have also aided in classification <strong>of</strong> the black soils as being mainly<br />

silty <strong>clays</strong> <strong>an</strong>d/ or silty <strong>clays</strong> with s<strong>an</strong>d, while varieties classified as clay <strong>an</strong>d clayey silt also<br />

occur but to a lesser extent. The red soils fall in the classes <strong>of</strong> clayey silts for depths <strong>of</strong> up to<br />

1,0m; <strong>an</strong>d silty <strong>clays</strong> with s<strong>an</strong>d, for depths greater th<strong>an</strong> 1,0m (Table 7.7). The classification<br />

implies that the fine fraction (silt <strong>an</strong>d clay sizes) predominates both types <strong>of</strong> <strong>clays</strong>, <strong>an</strong>d would<br />

most likely be the signific<strong>an</strong>t component in influencing <strong>an</strong>d/ or controlling the <strong>engineering</strong><br />

properties/ behaviour <strong>of</strong> these soils.<br />

9.3 Shear strength parameters<br />

The red soils are characterised by larger shear <strong>an</strong>gles (me<strong>an</strong> φ´: 28,5°) th<strong>an</strong> the black <strong>clays</strong><br />

(me<strong>an</strong> φ´: 18°); <strong>an</strong>d this implies that the former are relatively more stable <strong>an</strong>d have a larger<br />

capacity in supporting light <strong>engineering</strong> structures th<strong>an</strong> the latter.


178<br />

The variation <strong>of</strong> shear strength parameter φ´ for depths <strong>of</strong> less th<strong>an</strong> 0,5m (me<strong>an</strong> 21°, std 5,5)<br />

<strong>an</strong>d those greater th<strong>an</strong> 0,5m (me<strong>an</strong> 17°, std 3,8) in black <strong>clays</strong> was shown to be generally<br />

similar across the study area. This implies that geo-engineers <strong>an</strong>d pl<strong>an</strong>ners involved in<br />

pl<strong>an</strong>ning, design <strong>an</strong>d construction <strong>of</strong> projected structures could safely treat the black <strong>clays</strong> as a<br />

single homogeneous layer with respect to their strength characteristics.<br />

Results <strong>of</strong> correlations between index (Atterberg limits, relative consistency) <strong>an</strong>d shear<br />

strength (c´, φ´ ) parameters have been found to be only weak to moderate in strength. This<br />

implies that index parameters c<strong>an</strong>not be reliably used to characterise shear strength <strong>of</strong> soils in<br />

situ. The uncertainy accomp<strong>an</strong>ying the estimation <strong>an</strong>d assessment <strong>of</strong> strength characteristics<br />

could be a result <strong>of</strong> index tests having been performed on disturbed <strong>an</strong>d/ or fractioned samples<br />

having destroyed/ broken down soil fabric <strong>an</strong>d structure not representative <strong>of</strong> conditions in<br />

situ.<br />

The results <strong>of</strong> shear strength parameters (c´, φ´ ) obtained for black <strong>clays</strong> <strong>an</strong>d red soils in this<br />

study would serve to be useful in computations involving stability <strong>an</strong>alyses <strong>of</strong> soils when<br />

subjected to imposed loads from projected foundation structures. The parameters would<br />

especially be useful in determining the limiting value <strong>of</strong> external imposed loads capable <strong>of</strong><br />

just mobilising the shear strength <strong>of</strong> the soils, <strong>an</strong>d thereby cause their failure. In order to limit<br />

deformations to within tolerable limits, foundation design practices <strong>an</strong>d/ or procedures would<br />

require that a suitable factor <strong>of</strong> safety be adopted <strong>an</strong>d applied so that shear stresses induced in<br />

the soil due to weights <strong>of</strong> constructed structures are everywhere less th<strong>an</strong> a certain proportion<br />

<strong>of</strong> its maximum shear strength (Terzaghi <strong>an</strong>d Peck, 1967; Art. 26-32). The drained shear<br />

strength parameters would also serve to aid in the <strong>an</strong>alysis <strong>of</strong> long-term stability problems that<br />

may arise in a number <strong>of</strong> projects; whereupon the <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce would be useful in<br />

deriving earth pressure coefficients <strong>an</strong>d/ or bearing capacity coefficients for use in relev<strong>an</strong>t<br />

computations.<br />

The <strong>an</strong>alysis <strong>of</strong> long-term stability problems for the conditions prevailing after construction <strong>of</strong><br />

structures may include a number <strong>of</strong> determinations such as bearing capacity <strong>of</strong> footings <strong>an</strong>d<br />

foundations for structures on clay soils, earth pressure on a retaining wall, earth pressure<br />

against bracing in temporary excavations, safeguards against heave <strong>of</strong> the bottom <strong>of</strong><br />

temporary <strong>an</strong>d/ or deep open excavations in <strong>clays</strong>, as well as long-term stability <strong>of</strong> earth dams<br />

<strong>an</strong>d emb<strong>an</strong>kments (Terzaghi & Peck, 1967; Art. 29). In addition, residual shear strength<br />

parameters (cr´, φr´ ) when available, could be useful in long-term stability <strong>an</strong>alyses <strong>of</strong> slopes<br />

<strong>an</strong>d cuttings, especially in overconsolidated <strong>clays</strong> (Skempton, 1964; Skempton & La<br />

Rochelle, 1965; Symons, 1968). These parameters are usually obtained by allowing for large<br />

additional horizontal displacements during shearbox testing after the peak strength had been<br />

mobilised. However, their exact <strong>an</strong>d/ or accurate determination in this study was limited by<br />

the capacity <strong>of</strong> the apparatus. In practice, accurate determination <strong>of</strong> residual shear strength<br />

parameters is achieved by employing a ring shear apparatus which was, however, unavailable<br />

for the present study.<br />

9.4 Consolidation-settlement characteristics<br />

Results <strong>of</strong> laboratory consolidation tests (Tables 7.12 & 7.14 ) show that red soils exhibit<br />

relatively larger voids ratios <strong>an</strong>d higher porosity (initial voids ratio, eo: 1,56 at 0,30m depth &<br />

1,16 at 4m depth); with voids ratio, e, varying from 0,59-1,42 for the normal loading r<strong>an</strong>ge <strong>of</strong><br />

25-800 kPa. As a result <strong>of</strong> the above, together with their generally loose <strong>an</strong>d friable nature,<br />

these soils tend to be more or less free-draining, more permeable (K: 7,57E-11 to 7,33E-9<br />

m/s) <strong>an</strong>d more compressible (mv =0,16-2.94 m²/MN ). They would therefore exhibit faster <strong>an</strong>d


179<br />

rapid dissipation <strong>of</strong> pore water pressures <strong>an</strong>d faster consolidation settlements (cv = 1,18-12,14<br />

m²/year) when externally loaded. Constructed structures located on red soils are therefore<br />

expected to undergo faster consolidation –settlements, much <strong>of</strong> them during the construction<br />

stage, with minimum long-term instability implications after construction.<br />

On the other h<strong>an</strong>d, the black <strong>clays</strong> have been found to be characterised by relatively smaller<br />

voids ratios <strong>an</strong>d lower porosity (eo: 0,709-1,098; e = 0,23-1,06 over 25-800 kPa loading<br />

r<strong>an</strong>ge). These <strong>clays</strong> are generally more compact, dense <strong>an</strong>d cohesive, <strong>an</strong>d therefore less<br />

permeable (K = 1,67E-12 to 1,20E-9 m/s), less compressible (mv = 0,03-1,62 m²/MN) <strong>an</strong>d<br />

would exhibit slower dissipation <strong>of</strong> pore water pressures <strong>an</strong>d, therefore, slower consolidation<br />

settlements (cv = 0,05-5,08 m²/year) on external loading. The low rates <strong>of</strong> consolidationsettlement<br />

exhibited by the black <strong>clays</strong> would also me<strong>an</strong> structural settlements occurring<br />

beyond the construction stage. Long-term instability problems would therefore be expected<br />

for structures constructed on black <strong>clays</strong>.<br />

Results <strong>of</strong> oedometer consolidation tests obtained in this study could be used to solve<br />

problems related to foundations for light structures in the current area. The consolidation data<br />

<strong>an</strong>d parameters could be applied in conjuction with classification data, as well as knowledge<br />

<strong>of</strong> the loading history <strong>of</strong> the soils to make estimates <strong>an</strong>d predictions as regards probable<br />

behaviour <strong>of</strong> foundations under various loading conditions. The results would specifically<br />

assist in calculating the amount <strong>of</strong> settlement which would ultimately take place for the<br />

structure as a whole, while variations in long-term settlements <strong>an</strong>d/ or differential settlements<br />

between individual footings could also be estimated; by applying methods described by<br />

Terzaghi (1939), MacDonald <strong>an</strong>d Skempton (1955) <strong>an</strong>d Skempton <strong>an</strong>d Bjerum (1957).<br />

According to Skempton (1956), Terzaghi (1934), Mitchell, Vivatrat & Lambe (1977),<br />

differential settlements are usually more critical th<strong>an</strong> overall settlement, <strong>an</strong>d c<strong>an</strong> cause tilting<br />

<strong>of</strong> a structure as a whole <strong>an</strong>d/or distortions within the structure; <strong>an</strong>d should therefore be kept<br />

within acceptable limits to avoid possible structural deterioration <strong>an</strong>d damage. In addition,<br />

possible settlement <strong>of</strong> piled foundations as a result <strong>of</strong> the presence <strong>of</strong> deep-seated weak<br />

compressible clay layers, especially in red soils, could also be estimated.<br />

On the other h<strong>an</strong>d, estimates <strong>of</strong> the rate <strong>of</strong> consolidation could be further used to give the<br />

duration <strong>of</strong> time within which structural settlements would be completed, either during or<br />

after construction. In situations <strong>of</strong> long-term settlements as may occur in black <strong>clays</strong>,<br />

settlement/ time graphs could serve to show the duration <strong>of</strong> the most signific<strong>an</strong>t part <strong>of</strong><br />

settlements, <strong>an</strong>d this could be compared with the economic life <strong>of</strong> the structure. The<br />

settlement/ time relationship would also assist in ascertaining possible development <strong>of</strong><br />

unacceptable differential settlements in the long-term, after construction.<br />

Areas <strong>of</strong> s<strong>of</strong>t ground <strong>an</strong>d <strong>clays</strong> in the form <strong>of</strong> alluvial <strong>an</strong>d / or swampy environments are<br />

common in the present study area. The shear strength <strong>an</strong>d capacity <strong>of</strong> these grounds to support<br />

increased foundation loads could be increased <strong>an</strong>d improved by consolidation through preloading<br />

<strong>of</strong> the ground with a surcharge <strong>of</strong> temporary fill. In this case, results <strong>of</strong> laboratory<br />

consolidation tests would be useful in estimating the extent <strong>an</strong>d rate <strong>of</strong> the resulting<br />

settlement; <strong>an</strong>d thereby indicate whether a provision for accelerating the consolidation, such<br />

as installation <strong>of</strong> s<strong>an</strong>d drains, is justifiable.<br />

In situations where piles could be driven through the pre-consolidated s<strong>of</strong>t strata to tr<strong>an</strong>smit<br />

structural loads to a deeper firm stratum, the tendency <strong>of</strong> continued settlement <strong>of</strong> the fill <strong>an</strong>d<br />

s<strong>of</strong>t strata would have the effect <strong>of</strong> throwing additional loads onto the piles. Here also,


180<br />

knowledge <strong>of</strong> the consolidation characteristics would serve as a basis for safeguarding against<br />

overloading <strong>of</strong> the piles due to this effect.<br />

Results <strong>of</strong> consolidation tests could also be used to monitor ch<strong>an</strong>ges in effective strength in<br />

structures such as earth-dams <strong>an</strong>d emb<strong>an</strong>kments (which usually consolidate under their own<br />

weight); as well as in the underlying soil strata. This would ultimately serve to aid in the<br />

<strong>an</strong>alysis <strong>of</strong> the long-term stability <strong>of</strong> the structures.<br />

The use <strong>of</strong> results <strong>of</strong> oedometer consolidation tests is, however, usually limited when it comes<br />

to the estimation <strong>of</strong> rates <strong>of</strong> settlement <strong>of</strong> foundation soils. According to Rowe (1972), the<br />

underestimation is attributed to the small size <strong>of</strong> laboratory test specimens used which make it<br />

impracticable to represent m<strong>an</strong>y <strong>of</strong> the natural features <strong>of</strong> soil fabric (fissures, pores,<br />

laminations, other discontinuities), <strong>an</strong>d which have a pr<strong>of</strong>ound effect <strong>an</strong>d control on drainage<br />

<strong>an</strong>d therefore the rate <strong>of</strong> consolidation-settlement. As a result, a given proportion <strong>of</strong> the<br />

ultimate consolidation-settlement is reached in a shorter time in real field situations, th<strong>an</strong> that<br />

predicted from laboratory test data. In addition, extent <strong>of</strong> consolidation is based only on<br />

measurements <strong>of</strong> ch<strong>an</strong>ge in height <strong>of</strong> the specimen; while me<strong>an</strong>s for measuring excess pore<br />

pressure, the dissipation <strong>of</strong> which actually controls the consolidation process is, however, not<br />

provided for in the oedometer consolidation cell.<br />

9.5 Percentage swelling <strong>an</strong>d swelling pressure effects<br />

Results <strong>of</strong> swelling tests show that, in general, black <strong>clays</strong> would exhibit high percentage<br />

swelling (S > 75%) when imposed loads from light <strong>engineering</strong> structures located on <strong>an</strong>d/ or<br />

within, fall below 5 kPa. On the other h<strong>an</strong>d, external imposed loads greater th<strong>an</strong> 80 kPa would<br />

me<strong>an</strong> minimal swelling <strong>of</strong> the <strong>clays</strong>, i.e. S < 10% [S% in this case is the amount <strong>of</strong> swelling<br />

(S mm) under a given load decrement, expressed as a percentage <strong>of</strong> ultimate swelling (Smax<br />

mm) which occurs under zero external loading]. In addition, swelling pressure tests have<br />

shown the black <strong>clays</strong> to exhibit swelling pressures <strong>of</strong> 49-104 kPa. It would therefore be<br />

appropriate that in practice, light <strong>engineering</strong> structures are initially designed to impose loads<br />

<strong>of</strong> well over 100 kPa in order to minimise potential destabilisation effects that may arise from<br />

possible swelling <strong>of</strong> the <strong>clays</strong> on wetting.


Chapter 10<br />

181<br />

Correlation <strong>of</strong> index properties<br />

10.1 Linear shrinkage <strong>an</strong>d plasticity index<br />

Laboratory measured values <strong>of</strong> linear shrinkage, LS, <strong>an</strong>d plasticity index, PI meas.,<br />

Table 10.1. Calculated <strong>an</strong>d laboratory measured values <strong>of</strong> plasticity index <strong>of</strong> black <strong>clays</strong>.<br />

Pr<strong>of</strong>ile<br />

Depth<br />

(m)<br />

Sample No.<br />

LS<br />

(%)<br />

PI meas.<br />

(%) PI BS (%) PILS (%)<br />

Depth<br />

(m) Sample No. LS (%)<br />

PI meas.<br />

(%) PI BS (%)<br />

SA1-30cm 25 41 52 46 SA1-50cm 23 39 49 44<br />

A < 0,50 SA2-30cm 27 54 57 50 ≥ 0,50 SA2-70/105cm 26 55 55 49<br />

SA28-30cm 25 45 54 48 SA28-50cm 25 47 54 48<br />

SA37-30cm 24 43 51 45 SA37-50/80cm 26 39 55 49<br />

SA41-30cm 26 47 55 48 SA41-50cm 24 49 52 46<br />

SB1-30cm 23 47 49 44 SB1-50/70cm 25 50 53 47<br />

B < 0,50 SB5-30cm 23 47 49 44 ≥ 0,50 SB5-50/70cm 25 50 53 47<br />

SB7-30cm 24 48 51 45 SB7-50/70cm 25 49 53 47<br />

SB18-30cm 23 43 49 43 SB18-50cm 21 39 44 39<br />

SB27-30cm 24 47 50 44 SB27-50cm 25 49 52 46<br />

SB31-30cm 24 49 51 45 SB31-50cm 25 48 53 47<br />

SB41-30cm 26 44 55 49 SB41-50cm 25 47 54 48<br />

SB42-30cm 25 44 54 48 SB42-50cm 26 47 55 48<br />

SC1-30cm 23 48 49 43 SC1-50cm 24 48 51 45<br />

C < 0,50 SC5-30cm 23 48 49 44 ≥ 0,50 SC5-50cm 24 47 50 44<br />

SC9-30cm 24 48 50 44 SC9-50cm 24 48 51 45<br />

SC17-30cm 25 41 53 47 SC17-50cm 24 48 51 45<br />

SC25-30cm 25 49 52 46 SC25-50cm 25 48 54 48<br />

SC29-30cm 25 49 53 47 SC29-50cm 26 43 56 50<br />

SC33-30cm 25 49 53 47 SC33-50cm 26 50 55 48<br />

SC41-30cm 29 47 61 54 SC41-50cm 29 47 61 54<br />

SD1-30cm 26 46 56 49 SD1-50/80cm 27 48 57 50<br />

D < 0,50 SD5-30cm 26 47 55 48 ≥ 0,50 SD5-50/70cm 27 49 57 50<br />

SD7-30cm 27 48 57 50 SD7-50/70cm 26 51 56 50<br />

SD17-30cm 24 44 50 44 SD17-50cm 25 49 52 46<br />

SD25-30cm 25 39 53 47 SD25-50cm 25 45 52 46<br />

SD29-30cm 27 47 57 50 SD29-50cm 26 51 55 48<br />

SD33-30cm 27 47 58 51 SD33-50cm 25 50 53 47<br />

SD41-30cm 26 46 55 48 SD41-50cm 26 49 56 49<br />

SE1-30cm 23 48 49 44 SE1-50/70cm 25 50 53 47<br />

E < 0,50 SE5-30cm 24 48 50 44 ≥ 0,50 SE5-50/70cm 26 49 55 49<br />

SE13-30cm 24 47 51 45 SE13-50cm 25 49 53 47<br />

SE21-30cm 25 49 53 47 SE21-50cm 26 49 55 48<br />

SE29-30cm 25 46 52 46 SE29-50cm 26 47 56 50<br />

SE37-30cm 24 48 50 44 SE37-50cm 25 48 54 48<br />

SE41-30cm 25 47 53 47 SE41-50cm 26 48 55 48<br />

max 29 54 61 54 max 29 55 61 54<br />

Statistical < 0,50 min 23 39 49 43 ≥ 0,50 min 21 39 44 39<br />

<strong>an</strong>alysis r<strong>an</strong>ge 6 15 12 11 r<strong>an</strong>ge 8 16 17 15<br />

medi<strong>an</strong> 25 47 53 47 medi<strong>an</strong> 25 48 54 48<br />

mode 25 47 53 47 mode 25 49 53 47<br />

me<strong>an</strong> 25 47 53 47 me<strong>an</strong> 25 48 54 47<br />

std 1,38 2,78 2,93 2,59 std 1,29 3,28 2,76 2,43<br />

var 1,89 7,74 8,58 6,69 var 1,67 10,76 7,60 5,92<br />

cov 0,57 0,51 cov 3,23 2,85<br />

correlation 0,70 0,70 correlation 0,37 0,37<br />

PILS<br />

(%)


182<br />

Table 10.2. Calculated <strong>an</strong>d laboratory measured values <strong>of</strong> plasticity index <strong>of</strong> red soils.<br />

Soil type Depth (m) Sample No. LS (%) PI meas. (%) PIBS (%) PILS (%)<br />

0,30 Rd1-30cm 10 18 21 18<br />

Red soils 1,00 Rd1-100cm 11 21 23 20<br />

2,00 Rd1-200cm 11 20 23 20<br />

4,00 Rd1-400cm 11 18 23 20<br />

0,30 Rd2-30cm 10 19 21 18<br />

1,00 Rd2-100cm 11 22 23 20<br />

2,00 Rd2-200cm 11 21 23 20<br />

4,00 Rd2-400cm 11 19 23 20<br />

max 11 22 23 20<br />

Statistical min 10 18 21 18<br />

<strong>an</strong>alysis r<strong>an</strong>ge 1 5 2 2<br />

medi<strong>an</strong> 11 20 23 20<br />

mode 11 #NV 23 20<br />

me<strong>an</strong> 11 20 23 20<br />

std 0,46 1,62 0,99 0,85<br />

var 0,21 2,63 0,97 0,72<br />

cov 0,76 0,65<br />

correlation 0,54 0,54<br />

as obtained for the black <strong>clays</strong> in this study at depths <strong>of</strong> less th<strong>an</strong> 0,50m as well as those at<br />

0,50m <strong>an</strong>d greater, are given in Table (10.1).<br />

Calculated values <strong>of</strong> plasticity index, PI(BS), as estimated from laboratory determined linear<br />

shrinkage values using the British St<strong>an</strong>dard relationship (BS 1377: 1967), i.e.<br />

PI = 2,13 * LS ( derivation based on British soils) (10.1)<br />

are also included in the table for the two depth intervals. The calculated plasticity index<br />

values are observed to be overestimated with respect to the laboratory determined values for<br />

both the two depth intervals. A comparison <strong>of</strong> the respective maximum <strong>an</strong>d minimum values<br />

as well as the medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> calculated <strong>an</strong>d measured plasticity indices<br />

further indicates <strong>an</strong>d confirms <strong>an</strong> overestimation <strong>of</strong> the former with respect to the latter (Table<br />

10.1). The overestimation could be attributed to the generally higher org<strong>an</strong>ic content <strong>of</strong><br />

<strong>tropical</strong> black <strong>clays</strong> which tends to inhibit <strong>an</strong>d/ or suppress their plasticity to some degree.<br />

The coefficient <strong>of</strong> variation (covari<strong>an</strong>ce) with respect to the calculated <strong>an</strong>d measured values<br />

are 0,57 <strong>an</strong>d 3,23 for the two depth intervals, respectively. The st<strong>an</strong>dard deviation <strong>an</strong>d<br />

vari<strong>an</strong>ce for the calculated <strong>an</strong>d measured plasticity values are in close agreement <strong>an</strong>d more or<br />

less <strong>of</strong> the same order <strong>of</strong> magnitude in each case, implying a similar form <strong>an</strong>d/ or mode <strong>of</strong><br />

distribution <strong>an</strong>d variation across the study area.<br />

A correlation <strong>of</strong> laboratory measured values <strong>of</strong> plasticity index <strong>an</strong>d linear shrinkage is<br />

presented diagrammatically in Fig. (10.1).


183<br />

Black <strong>clays</strong>; plasticity index/ linear shrinkage<br />

60<br />

n = 36<br />

Plasticity index PI (%)<br />

55<br />

50<br />

45<br />

40<br />

35<br />

PI = 1,89LS<br />

PI = 1,87LS<br />

less th<strong>an</strong> 0,50m depth<br />

0,50m depth <strong>an</strong>d<br />

greater<br />

Linear (less th<strong>an</strong><br />

0,50m depth)<br />

Linear (0,50m depth<br />

<strong>an</strong>d greater)<br />

30<br />

20 25 30<br />

Linear shrinkage LS (%)<br />

Figure 10.1. Correlation <strong>of</strong> laboratory measured values <strong>of</strong> plasticity index <strong>an</strong>d linear<br />

shrinkage <strong>of</strong> black <strong>clays</strong>.<br />

From the diagram, the plasticity index <strong>of</strong> black <strong>clays</strong> could be approximately estimated by the<br />

relationship<br />

<strong>an</strong>d<br />

PI = 1,87*LS (10.2)<br />

PI = 1,89*LS (10.3)<br />

for soil depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong> depths <strong>of</strong> 0,50m <strong>an</strong>d greater, respectively. The<br />

two relationships derived for the two depth intervals are practically identical, <strong>an</strong>d this serves<br />

to support <strong>an</strong>d confirm <strong>an</strong> earlier suggestion that the black <strong>clays</strong> <strong>of</strong> the study area are<br />

homogeneous in their <strong>engineering</strong> character. On the average therefore, the plasticity index <strong>of</strong><br />

the whole soil thickness <strong>of</strong> black <strong>clays</strong> across the study area could be estimated from the<br />

average relationship <strong>of</strong><br />

PI = 1,88*LS (10.4)<br />

Calculated values <strong>of</strong> plasticity index, PILS, as estimated from laboratory measured values <strong>of</strong><br />

linear shrinkage, LS, using the above new relationship in Equation (10.4) are presented in<br />

Table (10.1), both for depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d greater. A<br />

comparison <strong>of</strong> the so calculated values <strong>of</strong> plasticity index (PILS) with plasticity index values<br />

measured in the laboratory (PI meas.) reveals a generally close agreement between the two sets<br />

<strong>of</strong> data. The maximum <strong>an</strong>d minimum values as well as the medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values<br />

obtained for the two sets <strong>of</strong> data are identical at each one <strong>of</strong> the two depth intervals. This<br />

implies that Equation (10.4) above is a better <strong>an</strong>d preferred estimator for the plasticity index<br />

<strong>of</strong> black <strong>clays</strong> th<strong>an</strong> the British st<strong>an</strong>dard relationship <strong>of</strong> Equation (10.1). In addition, values <strong>of</strong>


184<br />

st<strong>an</strong>dard deviation <strong>an</strong>d vari<strong>an</strong>ce derived for the calculated <strong>an</strong>d measured plasticity values<br />

(Table 10.1) are comparable <strong>an</strong>d <strong>of</strong> the same order <strong>of</strong> magnitude at each <strong>of</strong> the two depth<br />

intervals; <strong>an</strong>d this further pointing to the possible general uniform character <strong>an</strong>d/ or<br />

homogeneity <strong>of</strong> the soils across the study area. A covari<strong>an</strong>ce <strong>of</strong> 0,51 <strong>an</strong>d 2,85 is also<br />

registered for the two sets <strong>of</strong> data at depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d greater,<br />

respectively.<br />

A correlation <strong>of</strong> calculated <strong>an</strong>d laboratory measured plasticity indices is presented in Fig.<br />

(10.2).<br />

Black <strong>clays</strong>; measured/ calculated plasticity indices<br />

Measured plasticity index PImeas. (%)<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

n = 36<br />

PImeas = 1,01PILS<br />

(> 0,50m depth)<br />

PImeas. = 0,89PIBS.<br />

(> 0,50m depth)<br />

PImeas. = 1,00PILS<br />

(< 0,50m depth)<br />

PImeas. = 0,88PIBS<br />

(< 0,50m depth)<br />

PI (BS) vs PI (meas.), <<br />

0,50m depth<br />

PI (BS) vs PI (meas.),<br />

0,50m depth <strong>an</strong>d greater<br />

PI (LS) vs PI (meas.), <<br />

0,50m depth<br />

PI (LS) vs PI (meas.),<br />

0,50m depth <strong>an</strong>d greater<br />

30<br />

40 45 50 55 60 65<br />

Calculated plasticity index PILS (%)<br />

Figure 10.2. A correlation <strong>of</strong> calculated <strong>an</strong>d laboratory measured plasticity indices <strong>of</strong> black<br />

<strong>clays</strong>.<br />

It is observed that actual laboratory determined/ measured values <strong>of</strong> plasticity index (PImeas.)<br />

would usually be underestimated with respect to calculated plasticity values (PIBS) obtained<br />

using the British St<strong>an</strong>dard relationship (PI = 2,13*LS). However, the calculated <strong>an</strong>d measured<br />

plasticity indices are practically related in the same way for depths <strong>of</strong> less th<strong>an</strong> 0,50m (PImeas.<br />

= 0,88PIBS) <strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d greater (PImeas. = 0,89PIBS), implying homogeneous black<br />

<strong>clays</strong> across the study area. An average relationship <strong>of</strong><br />

PImeas. = 0,885PIBS (10.5)<br />

could therefore be safely adopted for the whole thickness <strong>of</strong> black <strong>clays</strong> across the study area.


185<br />

On the other h<strong>an</strong>d, laboratory measured plasticity values (PImeas.) compare exactly <strong>an</strong>d are<br />

practically the same as calculated values (PILS) derived using the new relationship (PI =<br />

1,88*LS) developed <strong>an</strong>d established in this study for the black <strong>clays</strong>. The relationship<br />

between the calculated <strong>an</strong>d measured plasticity indices is practically identical at both depths<br />

<strong>of</strong> less th<strong>an</strong> 0,50m i.e.<br />

PImeas. = 1,00PILS (10.6)<br />

<strong>an</strong>d those <strong>of</strong> 0,50m <strong>an</strong>d greater, i.e.<br />

PImeas. = 1,01PILS (10.7)<br />

so that the black <strong>clays</strong> must be uniform <strong>an</strong>d homogeneous in their <strong>engineering</strong> character<br />

across the study area. An average relationship <strong>of</strong><br />

PImeas. = 1,005PILS (10.8)<br />

could be adopted for the two sets <strong>of</strong> plasticity indices.<br />

The red soils involved in this study were investigated <strong>an</strong>d <strong>an</strong>alysed in a similar way.<br />

Laboratory determined values <strong>of</strong> plasticity index are presented alongside those derived by<br />

calculation using the British St<strong>an</strong>dard relationship in Table (10.2). A comparison <strong>of</strong> the<br />

minimum <strong>an</strong>d maximum values as well as the medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> the two sets<br />

<strong>of</strong> data reveals a slight overestimation <strong>of</strong> the calculated plasticity indices with respect to the<br />

laboratory measured values, <strong>an</strong>d this most probably due to reasons already stated above as for<br />

black <strong>clays</strong>. A covari<strong>an</strong>ce <strong>of</strong> 0,76 <strong>an</strong>d moderately strong correlation <strong>of</strong> 0,54 was also recorded<br />

for the two sets <strong>of</strong> data.<br />

A correlation <strong>of</strong> laboratory measured linear shrinkage (LS) <strong>an</strong>d plasticity index (PI) yielded<br />

new relationships for estimating the latter parameter from the former for the red soils (Fig.<br />

10.3), i.e.<br />

PI = 1,77LS (10.9)<br />

for the sampling area Rd1 <strong>an</strong>d<br />

PI = 1,90LS (10.10)<br />

for the sampling area Rd2, respectively. An average relationship <strong>of</strong><br />

PI = 1,83LS (10.11)<br />

was derived <strong>an</strong>d established for the red soils. Plasticity index values (PILS) derived by<br />

calculation from laboratory linear shrinkage (LS) values using this new relationship are also<br />

included in Table (10.2) for comparison purposes. The so calculated <strong>an</strong>d laboratory measured<br />

(PImeas.) plasticity index values are perfectly in close agreement <strong>an</strong>d/ or practically identical in<br />

terms <strong>of</strong> their maximum <strong>an</strong>d minimum values as well as medi<strong>an</strong> <strong>an</strong>d me<strong>an</strong> values. A<br />

covari<strong>an</strong>ce <strong>of</strong> 0,65 <strong>an</strong>d moderately strong correlation <strong>of</strong> 0,54 were registered for the two sets<br />

<strong>of</strong> data.


186<br />

Red soils; plasticity index/ linear shrinkage<br />

23<br />

22<br />

n = 8<br />

Plasticity index PI (%)<br />

21<br />

20<br />

19<br />

18<br />

17<br />

PI = 1,90LS<br />

PI = 1,77LS<br />

PI = 1,83LS<br />

Rd1+Rd2<br />

Rd1<br />

Rd2<br />

Linear (Rd1+Rd2)<br />

Linear (Rd1)<br />

Linear (Rd2)<br />

16<br />

15<br />

9,8 10 10,2 10,4 10,6 10,8 11 11,2<br />

Linear shrinkage LS (%)<br />

Figure 10.3. Correlation between laboratory measured values <strong>of</strong> plasticity index <strong>an</strong>d linear<br />

shrinkage <strong>of</strong> red soils.<br />

A correlation <strong>of</strong> laboratory measured plasticity indices (PImeas.) with plasticity indices derived<br />

by calculation (PIBS &PILS) using the British St<strong>an</strong>dard relationship as well as the new<br />

relationship in Equation (10.11) developed in this study for the red soils, is presented<br />

diagrammatically in Fig.( 10.4). Once again, laboratory measured plasticity indices would be<br />

slightly underestimated when derived from the British St<strong>an</strong>dard relationship, i.e.<br />

PImeas. = 0,86PIBS. (10.12)<br />

On the other h<strong>an</strong>d, the newly developed relationship (PI = 1,83LS) would perfectly reproduce<br />

<strong>an</strong>d/ or better estimate actual measured plasticity indices, with the measured <strong>an</strong>d calculated<br />

values being related in the form <strong>of</strong><br />

PImeas. = 1,00PILS (10.13)


187<br />

Red <strong>clays</strong>; measured/ calculated plasticity index<br />

23<br />

22<br />

n = 8<br />

Measured plasticitx index (%)<br />

21<br />

20<br />

19<br />

18<br />

17<br />

16<br />

PImeas. = 1,00PILS<br />

PImeas. = 0,86PIBS<br />

PI(BS) vs PI meas.<br />

PI (LS) vs PI meas.<br />

15<br />

15 17 19 21 23 25<br />

Calculated plasticity index (%)<br />

Figure 10.4. A correlation <strong>of</strong> calculated <strong>an</strong>d laboratory measured<br />

plasticity indices <strong>of</strong> red soils.<br />

10.2 Liquid limit <strong>an</strong>d plasticity index<br />

The plasticity index <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils involved in this study could be estimated by<br />

calculation from laboratory determined results <strong>of</strong> liquid limit using the British St<strong>an</strong>dard<br />

relationship<br />

PI =0,73(LL-20) (10.14)<br />

through which they are very strongly interrelated, i.e.R = 1 (Fig. 10.5). A correlation <strong>of</strong><br />

laboratory determined values <strong>of</strong> plasticity index <strong>an</strong>d liquid limit have also given rise to a new<br />

<strong>an</strong>d strong (R = 0,96) relationship, i.e.<br />

PI = 0,79(LL-25) (10.15)<br />

Diagram: plasticity index/ liquid limit<br />

Plasticity index PI (%)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

n = 90<br />

PIBS = 0,73(LL - 20)<br />

R 2 = 1<br />

PI = 0,79(LL - 25)<br />

R 2 = 0,92<br />

40 60 80 100<br />

Liquid limit LL (%)<br />

PI British Std<br />

Black+red soils<br />

Fig. 10.5. Correlation: laboratory measured values <strong>of</strong> plasticity index<br />

<strong>an</strong>d liquid limit <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.


188<br />

Table 10.3. Calculated <strong>an</strong>d laboratory measured plasticity<br />

indices <strong>of</strong> black <strong>clays</strong>.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%)<br />

PI<br />

meas. PI BS<br />

(%) (%) PILL (%)<br />

SA1-30cm 32 78 37 41 42 42<br />

A SA1-50cm 30 75 36 39 40 39<br />

SA2-30cm 34 93 39 54 53 53<br />

(black <strong>clays</strong>) SA2-70cm 34 90 36 54 51 51<br />

SA2-105cm 30 88 33 55 50 50<br />

SA28-30cm 26 85 40 45 47 47<br />

SA28-50cm 27 89 41 47 50 50<br />

SA37-30cm 19 81 38 43 44 44<br />

SA37-50cm 22 83 39 44 46 46<br />

SA37-80cm 23 85 39 46 48 47<br />

SA41-30cm 23 89 41 47 50 50<br />

SA41-50cm 25 90 41 49 51 51<br />

SB1-30cm 23 82 34 47 45 45<br />

B SB1-50cm 27 84 36 48 47 46<br />

SB1-70cm 30 90 39 51 51 51<br />

(black <strong>clays</strong>) SB5-30cm 24 81 34 47 44 44<br />

SB5-50cm 27 85 37 48 47 47<br />

SB5-70cm 31 91 40 51 52 52<br />

SB7-30cm 23 83 35 48 46 46<br />

SB7-50cm 26 86 37 49 48 48<br />

SB7-70cm 30 87 38 49 49 49<br />

SB18-30cm 20 82 40 43 46 45<br />

SB18-50cm 20 72 34 39 38 37<br />

SB27-30cm 22 85 38 47 47 47<br />

SB27-50cm 22 88 39 49 49 50<br />

SB31-30cm 23 87 38 49 49 49<br />

SB31-50cm 23 86 38 48 48 48<br />

SB41-30cm 23 89 45 44 50 50<br />

SB41-50cm 24 91 44 47 52 52<br />

SB42-30cm 22 88 44 44 49 49<br />

SB42-50cm 21 91 45 47 52 52<br />

SC1-30cm 24 82 35 48 45 45<br />

C SC1-50cm 26 85 37 48 47 47<br />

SC5-30cm 23 83 35 48 46 46<br />

(black <strong>clays</strong>) SC5-50cm 25 84 36 47 46 46<br />

SC9-30cm 25 84 36 48 46 46<br />

SC9-50cm 26 85 37 48 47 47<br />

SC17-30cm 21 76 35 41 41 40<br />

SC17-50cm 22 81 33 48 44 44<br />

SC25-30cm 26 88 39 49 50 50<br />

SC25-50cm 26 89 41 48 50 50<br />

SC29-30cm 21 82 32 49 45 45<br />

SC29-50cm 20 80 37 43 44 44<br />

SC33-30cm 20 82 33 49 45 45<br />

SC33-50cm 21 83 33 50 46 46<br />

SC41-30cm 18 85 38 47 48 48<br />

SC41-50cm 17 84 37 47 46 46


189<br />

These two relationships could be especially useful in the approximate estimation <strong>of</strong> the<br />

plasticity index from measured values <strong>of</strong> liquid limit for the more silty varieties <strong>of</strong> clay soils<br />

in which determination <strong>of</strong> the plastic limit could otherwise be difficult.<br />

Table 10.3, continued: Calculated <strong>an</strong>d laboratory measured plasticity indices <strong>of</strong> black <strong>clays</strong>.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%) PI meas(%) PI BS (%) PILL (%)<br />

SD1-30cm 27 82 36 46 45 45<br />

D SD1-50cm 27 83 36 47 46 46<br />

SD1-80cm 30 91 42 49 52 52<br />

(black <strong>clays</strong>) SD5-30cm 26 84 36 47 46 46<br />

SD5-50cm 27 84 36 48 47 47<br />

SD5-80cm 30 92 42 50 53 53<br />

SD7-30cm 27 84 36 48 47 47<br />

SD7-50cm 27 86 37 49 48 48<br />

SD7-70cm 30 91 39 52 52 52<br />

SD17-30cm 20 84 40 44 47 46<br />

SD17-50cm 22 84 35 49 47 47<br />

SD25-30cm 22 80 41 39 44 44<br />

SD25-50cm 21 84 40 45 47 47<br />

SD29-30cm 18 85 38 47 48 48<br />

SD29-50cm 22 86 35 51 48 48<br />

SD33-30cm 20 85 37 47 47 47<br />

SD33-50cm 21 86 35 50 48 48<br />

SD41-30cm 23 85 39 46 47 47<br />

SD41-50cm 25 90 41 49 51 51<br />

SE1-30cm 24 82 35 48 45 45<br />

E SE1-50cm 27 85 37 48 47 47<br />

SE1-70cm 30 91 39 52 52 52<br />

(black <strong>clays</strong>) SE5-30cm 25 85 37 48 47 47<br />

SE5-50cm 27 86 38 48 48 48<br />

SE5-70cm 29 92 42 50 53 53<br />

SE13-30cm 21 86 39 47 48 48<br />

SE13-50cm 24 89 40 49 50 50<br />

SE21-30cm 25 88 39 49 50 50<br />

SE21-50cm 26 90 41 49 51 51<br />

SE29-30cm 22 81 36 46 44 44<br />

SE29-50cm 23 83 36 47 46 45<br />

SE37-30cm 21 85 37 48 47 47<br />

SE37-50cm 22 85 38 48 48 48<br />

SE41-30cm 22 87 40 47 49 49<br />

SE41-50cm 21 88 40 48 49 49<br />

max 55 53 53<br />

Statistical min 39 38 37<br />

<strong>an</strong>alysis r<strong>an</strong>ge 17 15 16<br />

medi<strong>an</strong> 48 47 47<br />

mode 48 47 47<br />

me<strong>an</strong> 47 48 48<br />

st<strong>an</strong>dard dev. 3,00 2,85 3,09<br />

vari<strong>an</strong>ce 9,00 8,13 9,53<br />

covari<strong>an</strong>ce 5,89 6,37<br />

correlation 0,70 0,70<br />

Results <strong>of</strong> plasticity index values, PIBS <strong>an</strong>d PILL, derived by calculation from laboratory<br />

measured liquid limits using the British St<strong>an</strong>dard relationship (Equation 10.14) <strong>an</strong>d the new<br />

relationship developed in this study (Equation 10.15), respectively, are presented in Table (


190<br />

10.3) for the black <strong>clays</strong>. The calculated values are shown to be strongly in close agreement<br />

with laboratory determined plasticity indices (PImeasured) in terms <strong>of</strong> maximum (53-55%),<br />

minimum (37-39%) <strong>an</strong>d r<strong>an</strong>ge (15-17%) values; as well as the medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong><br />

values (47-48%). St<strong>an</strong>dard deviation (2,85-3,09) <strong>an</strong>d vari<strong>an</strong>ce (8,13-9,53) are also in strong<br />

agreement <strong>an</strong>d <strong>of</strong> the same order <strong>of</strong> magnitude, signifying a similar form <strong>of</strong> distribution across<br />

the study area. A covari<strong>an</strong>ce <strong>of</strong> 5,89-6,37 <strong>an</strong>d a strong correlation <strong>of</strong> 0,70 was also registered<br />

between the calculated <strong>an</strong>d laboratory determined plasticity indices.<br />

Table 10.4. Calculated <strong>an</strong>d laboratory measured plasticity indices <strong>of</strong> red soils.<br />

Pr<strong>of</strong>ile Sample No. Wn (%) LL (%) PL (%)<br />

PImeasured<br />

(%) PI BS (%) PILL (%)<br />

Rd1 -30cm 39 49 31 18 21 19<br />

Red soils Rd1-100cm 24 51 31 21 23 21<br />

Rd1-200cm 24 49 30 20 21 19<br />

Rd1-400cm 25 48 30 18 20 18<br />

Rd2-30cm 38 53 35 19 24 22<br />

Rd2-100cm 26 57 34 22 27 25<br />

Rd2-200cm 25 54 33 21 25 23<br />

Rd2-400cm 26 53 34 19 24 22<br />

max 22 27 25<br />

Statistical min 18 20 18<br />

<strong>an</strong>alysis r<strong>an</strong>ge 5 7 7<br />

medi<strong>an</strong> 20 24 21<br />

mode #NV 24 22<br />

me<strong>an</strong> 20 23 21<br />

st<strong>an</strong>dard dev. 1,62 2,29 2,48<br />

vari<strong>an</strong>ce 2,63 5,26 6,16<br />

covari<strong>an</strong>ce 2,56 2,77<br />

correlation 0,79 0,79<br />

Calculated plasticity indices for red soils (Table 10.4) are also closely comparable to the<br />

laboratory determined plasticity indices, with maximum <strong>an</strong>d minimum values <strong>of</strong> 22-27% <strong>an</strong>d<br />

18-20%, as well as a r<strong>an</strong>ge <strong>of</strong> 5-7%. Medi<strong>an</strong>, mode, <strong>an</strong>d me<strong>an</strong> values <strong>of</strong> 20-24%, 22-24% <strong>an</strong>d<br />

20-23% were also recorded. The st<strong>an</strong>dard deviation <strong>an</strong>d vari<strong>an</strong>ce are <strong>of</strong> the same order <strong>of</strong><br />

magnitude while a covari<strong>an</strong>ce <strong>of</strong> 2,56-2,77 <strong>an</strong>d strong correlation <strong>of</strong> 0,79 was recorded<br />

between the calculated <strong>an</strong>d measured sets <strong>of</strong> data.<br />

The calculated <strong>an</strong>d laboratory measured plasticity indices for both black <strong>clays</strong> <strong>an</strong>d red soils<br />

are in a very strong <strong>an</strong>d near perfect agreement (Fig. 10.6), being related in the form <strong>of</strong><br />

<strong>an</strong>d<br />

PImeasured = 0,99PIBS, (10.16)<br />

PImeasured = 0,99PILL (10.17)


191<br />

Diagram: PImeasured/ PIcalculated<br />

60<br />

Measured plasticity index (%)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

n = 90<br />

PImeasured = 0,99PILL<br />

R 2 = 0,92<br />

PImeasured = 0,99PIBS<br />

R 2 = 0,92<br />

PI( BS) vs PI (measured)<br />

PI (LL) vs PI (measured)<br />

0<br />

0 20 40 60<br />

Calculated plasticity index (%)<br />

Figure 10.6. A correlation between calculated <strong>an</strong>d laboratory determined plasticity<br />

indices <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

10.3 Liquid limit <strong>an</strong>d Linear shrinkage<br />

A correlation <strong>of</strong> laboratory determined values <strong>of</strong> linear shrinkage (LS) <strong>an</strong>d liquid limit (LL)<br />

results in a very strong combined relationship (R = 0,93) for both black <strong>clays</strong> <strong>an</strong>d red soils as<br />

illustrated in Fig. (10.7). The relationship could be summarised in the form <strong>of</strong> <strong>an</strong> equation, i.e.<br />

LS = 0,39(LL-21) (10.18)<br />

Diagram: linear shrinkage/ liquid limit<br />

Linear shrinkage LS (%)<br />

35<br />

n = 90<br />

30<br />

LS = 0,39(LL - 21)<br />

25<br />

R 2 = 0,87<br />

20<br />

15<br />

10<br />

5<br />

0<br />

40 50 60 70 80 90 100<br />

Liquid limit LL (%)<br />

Black <strong>clays</strong><br />

+ red soils<br />

Figure 10.7. Correlation between laboratory measured values <strong>of</strong> linear shrinkage<br />

<strong>an</strong>d liquid limit <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.


192<br />

Calculated values <strong>of</strong> linear shrinkage computed from laboratory liquid limit values using this<br />

new relationship are presented in Table (10.5).<br />

Table 10.5. Calculated <strong>an</strong>d laboratory measured linear shrinkage<br />

values <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Pr<strong>of</strong>ile Sample No. LL (%) LS measured (%) LSLL (%)<br />

SA1-30cm 78 25 22<br />

A SA1-50cm 75 23 21<br />

SA2-30cm 93 27 28<br />

(black <strong>clays</strong>) SA2-70cm 90 25 27<br />

SA2-105cm 88 26 26<br />

SA28-30cm 85 25 25<br />

SA28-50cm 89 25 26<br />

SA37-30cm 81 24 23<br />

SA37-50cm 83 26 24<br />

SA37-80cm 85 26 25<br />

SA41-30cm 89 26 26<br />

SA41-50cm 90 24 27<br />

SB1-30cm 82 23 24<br />

B SB1-50cm 84 24 24<br />

SB1-70cm 90 25 27<br />

(black <strong>clays</strong>) SB5-30cm 81 23 23<br />

SB5-50cm 85 24 25<br />

SB5-70cm 91 26 27<br />

SB7-30cm 83 24 24<br />

SB7-50cm 86 24 25<br />

SB7-70cm 87 25 26<br />

SB18-30cm 82 23 24<br />

SB18-50cm 72 21 20<br />

SB27-30cm 85 24 25<br />

SB27-50cm 88 25 26<br />

SB31-30cm 87 24 26<br />

SB31-50cm 86 25 25<br />

SB41-30cm 89 26 26<br />

SB41-50cm 91 25 27<br />

SB42-30cm 88 25 26<br />

SB42-50cm 91 26 27<br />

SC1-30cm 82 23 24<br />

C SC1-50cm 85 24 25<br />

SC5-30cm 83 23 24<br />

(black <strong>clays</strong>) SC5-50cm 84 24 24<br />

SC9-30cm 84 24 24<br />

SC9-50cm 85 24 25<br />

SC17-30cm 76 25 21<br />

SC17-50cm 81 24 23<br />

SC25-30cm 88 25 26<br />

SC25-50cm 89 25 26<br />

SC29-30cm 82 25 24<br />

SC29-50cm 80 26 23<br />

SC33-30cm 82 25 24<br />

SC33-50cm 83 26 24<br />

SC41-30cm 85 29 25<br />

SC41-50cm 84 29 24


193<br />

Table 10.5, continued. Calculated <strong>an</strong>d laboratory measured linear<br />

shrinkage values <strong>of</strong> black <strong>an</strong>d red soils.<br />

Pr<strong>of</strong>ile Sample No. LL (%) LS measured (%) LS calculated (%)<br />

SD1-30cm 82 26 24<br />

D SD1-50cm 83 27 24<br />

SD1-80cm 91 26 27<br />

(black <strong>clays</strong>) SD5-30cm 84 26 24<br />

SD5-50cm 84 27 25<br />

SD5-80cm 92 27 28<br />

SD7-30cm 84 27 25<br />

SD7-50cm 86 26 25<br />

SD7-70cm 91 26 27<br />

SD17-30cm 84 24 24<br />

SD17-50cm 84 25 25<br />

SD25-30cm 80 25 23<br />

SD25-50cm 84 25 25<br />

SD29-30cm 85 27 25<br />

SD29-50cm 86 26 25<br />

SD33-30cm 85 27 25<br />

SD33-50cm 86 25 25<br />

SD41-30cm 85 26 25<br />

SD41-50cm 90 26 27<br />

SE1-30cm 82 23 24<br />

E SE1-50cm 85 24 25<br />

SE1-70cm 91 25 27<br />

(black <strong>clays</strong>) SE5-30cm 85 24 25<br />

SE5-50cm 86 25 25<br />

SE5-70cm 92 26 28<br />

SE13-30cm 86 24 25<br />

SE13-50cm 89 25 26<br />

SE21-30cm 88 25 26<br />

SE21-50cm 90 26 27<br />

SE29-30cm 81 25 23<br />

SE29-50cm 83 26 24<br />

SE37-30cm 85 24 25<br />

SE37-50cm 85 25 25<br />

SE41-30cm 87 25 26<br />

SE41-50cm 88 26 26<br />

Rd1 -30cm 49 10 11<br />

Red soils Rd1-100cm 51 11 12<br />

Rd1-200cm 49 11 11<br />

Rd1-400cm 48 11 10<br />

Rd2-30cm 53 10 13<br />

Rd2-100cm 57 11 14<br />

Rd2-200cm 54 11 13<br />

Rd2-400cm 53 11 13<br />

max 29 28<br />

min 10 10<br />

r<strong>an</strong>ge 19 18<br />

medi<strong>an</strong> 25 25<br />

mode 25 25<br />

me<strong>an</strong> 24 24<br />

st<strong>an</strong>dard dev. 4,30 4,02<br />

covari<strong>an</strong>ce 15,88<br />

correlation 0,93


194<br />

A comparison <strong>of</strong> laboratory determined linear shrinkage values (LS measured) with those<br />

computed (LSLL) according to Equation (10.18) reveals a very strong closeness. The two sets<br />

<strong>of</strong> data are practically identical in terms <strong>of</strong> maximun <strong>an</strong>d minimum values, the r<strong>an</strong>ge as well<br />

as the medi<strong>an</strong>, mode <strong>an</strong>d me<strong>an</strong> values (Table 10.5). The two comparable st<strong>an</strong>dard deviations<br />

<strong>of</strong> 4,30 <strong>an</strong>d 4,02, <strong>an</strong>d very strong correlation <strong>of</strong> 0,93 are further evidence <strong>of</strong> the similarity <strong>of</strong><br />

the two sets <strong>of</strong> data; <strong>an</strong>d therefore the reliability <strong>of</strong> the new relationship in estimating linear<br />

shrinkage <strong>of</strong> soils from laboratory measured liquid limits. The relationship between laboratory<br />

measured <strong>an</strong>d calculated linear shrinkage values is illustrated diagrammatically in Fig. (10.8),<br />

<strong>an</strong>d takes the form <strong>of</strong><br />

LSmeasured = 0,997LSLL (10.19)<br />

LS measured/ LSLL<br />

Measured linear shrinkage<br />

LSmeas. (%)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

n = 90<br />

R 2 = 0,86<br />

LS meas. = 0,997LSLL<br />

Black <strong>clays</strong> <strong>an</strong>d red<br />

soils<br />

Linear (Black <strong>clays</strong><br />

<strong>an</strong>d red soils)<br />

0 10 20 30<br />

Calculated linear shrinkage LSLL<br />

(%)<br />

Figure 10.8. A correlation between calculated <strong>an</strong>d laboratory determined linear<br />

shrinkage <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

10.4 Free swell <strong>an</strong>d other index properties<br />

The Swelling capability <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils could be estimated with a high degree <strong>of</strong><br />

approximation (R² = 0,85 <strong>an</strong>d over) from laboratory results <strong>of</strong> Atterberg limits <strong>an</strong>d linear<br />

shrinkage. The form <strong>an</strong>d strength <strong>of</strong> correlation between laboratory determined free swell<br />

(FS) on one h<strong>an</strong>d; <strong>an</strong>d liquid limits (LL), plasticity index (PI) <strong>an</strong>d linear shrinkage (LS) <strong>of</strong> the<br />

soils on the other, are illustrated in Table (10.6) <strong>an</strong>d Fig. (10.9).


195<br />

Table 10.6. Calculated <strong>an</strong>d laboratory measured free swell values <strong>of</strong> black<br />

<strong>clays</strong> <strong>an</strong>d red soils.<br />

Pr<strong>of</strong>ile Sample No. FS (%) LL (%) PL (%) PI (%) LS (%) FSLL (%) FSPI (%) FSLS (%)<br />

SA1-30cm 135 78 37 41 25 105 106 131<br />

A SA1-50cm 100 75 36 39 23 95 97 117<br />

SA2-30cm 135 93 39 54 27 151 155 146<br />

(black <strong>clays</strong>) SA2-70cm 170 90 36 54 25 143 154 131<br />

SA2-105cm 120 88 33 55 26 137 160 139<br />

SA28-30cm 125 85 40 45 25 126 122 131<br />

SA28-50cm 135 89 41 47 25 138 130 131<br />

SA37-30cm 115 81 38 43 24 114 113 124<br />

SA37-50cm 115 83 39 44 26 121 118 139<br />

SA37-80cm 115 85 39 46 26 127 125 139<br />

SA41-30cm 135 89 41 47 26 138 130 139<br />

SA41-50cm 120 90 41 49 24 142 135 124<br />

SB1-30cm 125 82 34 47 23 116 130 117<br />

B SB1-50cm 135 84 36 48 24 123 131 124<br />

SB1-70cm 145 90 39 51 25 142 143 131<br />

(black <strong>clays</strong>) SB5-30cm 120 81 34 47 23 114 129 117<br />

SB5-50cm 135 85 37 48 24 126 133 124<br />

SB5-70cm 145 91 40 51 26 146 144 139<br />

SB7-30cm 130 83 35 48 24 120 134 124<br />

SB7-50cm 135 86 37 49 24 129 136 124<br />

SB7-70cm 140 87 38 49 25 133 135 131<br />

SB18-30cm 120 82 40 43 23 119 112 117<br />

SB18-50cm 120 72 34 39 21 87 98 102<br />

SB27-30cm 130 85 38 47 24 127 129 124<br />

SB27-50cm 130 88 39 49 25 135 135 131<br />

SB31-30cm 130 87 38 49 24 133 136 124<br />

SB31-50cm 130 86 38 48 25 130 134 131<br />

SB41-30cm 145 89 45 44 26 139 118 139<br />

SB41-50cm 145 91 44 47 25 145 129 131<br />

SB42-30cm 145 88 44 44 25 135 117 131<br />

SB42-50cm 145 91 45 47 26 146 127 139<br />

SC1-30cm 125 82 35 48 23 117 131 117<br />

C SC1-50cm 135 85 37 48 24 126 133 124<br />

SC5-30cm 125 83 35 48 23 121 134 117<br />

(black <strong>clays</strong>) SC5-50cm 135 84 36 47 24 122 130 124<br />

SC9-30cm 130 84 36 48 24 122 133 124<br />

SC9-50cm 135 85 37 48 24 126 131 124<br />

SC17-30cm 105 76 35 41 25 97 106 131<br />

SC17-50cm 110 81 33 48 24 113 133 124<br />

SC25-30cm 125 88 39 49 25 136 136 131<br />

SC25-50cm 145 89 41 48 25 138 131 131<br />

SC29-30cm 110 82 32 49 25 116 138 131<br />

SC29-50cm 110 80 37 43 26 112 115 139<br />

SC33-30cm 120 82 33 49 25 116 137 131<br />

SC33-50cm 120 83 33 50 26 120 140 139<br />

SC41-30cm 125 85 38 47 29 127 130 160<br />

SC41-50cm 125 84 37 47 29 122 129 160


196<br />

Table 10.6, continued. Calculated <strong>an</strong>d laboratory measured free swell values<br />

<strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Pr<strong>of</strong>ile Sample No. FS (%) LL (%) PL (%) PI (%) LS (%) FSLL (%) FSPI (%) FSLS (%)<br />

SD1-30cm 125 82 36 46 26 118 127 139<br />

D SD1-50cm 130 83 36 47 27 120 128 146<br />

SD1-80cm 140 91 42 49 26 146 138 139<br />

(black <strong>clays</strong>) SD5-30cm 125 84 36 47 26 122 129 139<br />

SD5-50cm 130 84 36 48 27 124 132 146<br />

SD5-80cm 140 92 42 50 27 150 142 146<br />

SD7-30cm 110 84 36 48 27 125 133 146<br />

SD7-50cm 140 86 37 49 26 130 136 139<br />

SD7-70cm 135 91 39 52 26 145 148 139<br />

SD17-30cm 120 84 40 44 24 123 117 124<br />

SD17-50cm 130 84 35 49 25 123 138 131<br />

SD25-30cm 125 80 41 39 25 111 98 131<br />

SD25-50cm 135 84 40 45 25 125 120 131<br />

SD29-30cm 125 85 38 47 27 128 130 146<br />

SD29-50cm 145 86 35 51 26 130 143 139<br />

SD33-30cm 125 85 37 47 27 126 130 146<br />

SD33-50cm 145 86 35 50 25 129 141 131<br />

SD41-30cm 130 85 39 46 26 126 124 139<br />

SD41-50cm 135 90 41 49 26 143 136 139<br />

SE1-30cm 125 82 35 48 23 118 131 117<br />

E SE1-50cm 135 85 37 48 24 126 131 124<br />

SE1-70cm 145 91 39 52 25 145 146 131<br />

(black <strong>clays</strong>) SE5-30cm 125 85 37 48 24 126 132 124<br />

SE5-50cm 130 86 38 48 25 128 131 131<br />

SE5-70cm 145 92 42 50 26 150 140 139<br />

SE13-30cm 125 86 39 47 24 129 128 124<br />

SE13-50cm 130 89 40 49 25 138 135 131<br />

SE21-30cm 125 88 39 49 25 137 137 131<br />

SE21-50cm 145 90 41 49 26 143 138 139<br />

SE29-30cm 120 81 36 46 25 114 123 131<br />

SE29-50cm 120 83 36 47 26 119 127 139<br />

SE37-30cm 125 85 37 48 24 126 131 124<br />

SE37-50cm 130 85 38 48 25 128 131 131<br />

SE41-30cm 135 87 40 47 25 132 128 131<br />

SE41-50cm 135 88 40 48 26 135 131 139<br />

Rd1 -30cm 20 49 31 18 10 11 19 22<br />

Red soils Rd1-100cm 15 51 31 21 11 19 28 29<br />

Rd1-200cm 20 49 30 20 11 14 26 29<br />

Rd1-400cm 15 48 30 18 11 9 18 29<br />

Rd2-30cm 20 53 35 19 10 26 21 22<br />

Rd2-100cm 15 57 34 22 11 37 36 29<br />

Rd2-200cm 15 54 33 21 11 29 31 29<br />

Rd2-400cm 15 53 34 19 11 26 24 29<br />

max 170 151 160 160<br />

min 15 9 18 22<br />

Statistical r<strong>an</strong>ge 155 142 142 139<br />

<strong>an</strong>alysis medi<strong>an</strong> 130 126 131 131<br />

mode 125 126 131 131<br />

me<strong>an</strong> 120 118 121 123<br />

st<strong>an</strong>dard dev. 34 33 32 31<br />

covari<strong>an</strong>ce 1053 1013 970<br />

correlation 0,96 0,94 0,92


197<br />

The swelling capability in terms <strong>of</strong> free swell (FS) could be estimated from the Atterberg<br />

limits <strong>an</strong>d linear shrinkage <strong>an</strong>d vice versa, using the relationships,<br />

<strong>an</strong>d<br />

FS = 3,17(LL-45) (10.20)<br />

FS = 3,79(PI-13) (10.21)<br />

FS = 7,29(LS-7) (10.22)<br />

Free swell/ other index properties<br />

180<br />

160<br />

140<br />

FS = 7,29(LS - 7)<br />

R 2 = 0,85<br />

n = 90<br />

FS = 3,17(LL - 45)<br />

R 2 = 0,92<br />

Free swell FS (%)<br />

120<br />

100<br />

80<br />

60<br />

FS = 3,79(PI - 13)<br />

R 2 = 0,88<br />

FS vs LL<br />

FS vs PI<br />

FS vs LS<br />

40<br />

20<br />

0<br />

0 20 40 60 80 100<br />

Liquid limit LL, plasticity index PI <strong>an</strong>d linear<br />

shrinkage LS (%)<br />

Figure 10.9. Correlation: laboratory measured values <strong>of</strong> free swell <strong>an</strong>d laboratory<br />

liquid limit, plasticity index <strong>an</strong>d linear shrinkage <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

The strength <strong>of</strong> agreement between laboratory measured free swell values (FS meas.) <strong>an</strong>d those<br />

calculated by computation from laboratory Atterberg limits(FSLL; FSPI) <strong>an</strong>d linear shrinkage<br />

(FSLS) using the above new relationships is very high (Table 10.6); as characterised by very<br />

strong correlations <strong>of</strong> 0,92 <strong>an</strong>d over. Further evidence in their closeness is revealed by the<br />

medi<strong>an</strong> (126-131%), mode (125-131%), me<strong>an</strong> (118-123%) <strong>an</strong>d st<strong>an</strong>dard deviation (31-34)<br />

values which are found to be closely comparable <strong>an</strong>d <strong>of</strong> the same order <strong>of</strong> magnitude.<br />

The relationships between laboratory measured <strong>an</strong>d calculated free swell values (Fig. 10.10)<br />

take the form <strong>of</strong><br />

FSmeasured = 1,01FSLL (10.23)


198<br />

FSmeasured = 0,99FSPI (10.24)<br />

<strong>an</strong>d<br />

FSmeasured = 0,98FSLS (10.25)<br />

These relationships serve to support the reliability with which laboratory free swell; <strong>an</strong>d<br />

therefore swelling capability <strong>of</strong> soils could be estimated from laboratory determined Atterberg<br />

limits <strong>an</strong>d linear shrinkage.<br />

Measured/ calculated free swell<br />

Measured free swell (%)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

n = 91<br />

FSmeas. = 0,99FSPI<br />

FSmeas. = 1,01FSLL<br />

FSmeas. = 0,98FSLS<br />

0 50 100 150 200<br />

FS (meas.) vs FS (LL)<br />

FS (meas.) vs FS (PI)<br />

FS (meas.) vs FS (LS)<br />

Calculated free swell (%)<br />

Figure 10.10. A correlation between calculated <strong>an</strong>d laboratory determined free<br />

swell values <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

10.5 Clay fraction <strong>an</strong>d linear shrinkage, free swell<br />

Relatioships between clay fraction <strong>an</strong>d linear shrinkage <strong>an</strong>d/ or swelling capability <strong>of</strong> black<br />

<strong>clays</strong> <strong>an</strong>d red soils are generally poorly defined, being characterised by very weak correlations<br />

<strong>of</strong> less th<strong>an</strong> 0,3 (Figs. 10.11 <strong>an</strong>d 10.12).<br />

Linear shrinkage/ clay fraction<br />

Linear shrinkage LS (%)<br />

35<br />

30<br />

n = 87<br />

25<br />

20<br />

15<br />

R 2 = 0,0363<br />

10<br />

5<br />

0<br />

10 20 30 40 50<br />

% clay<br />

LS vs % clay<br />

Figure 10.11. Correlation between laboratory measured values <strong>of</strong> linear<br />

shrinkage <strong>an</strong>d clay fraction <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.


199<br />

Free swell/ clay fraction<br />

Free swell FS (%)<br />

200<br />

n = 87<br />

150<br />

100<br />

R 2 = 0,0207<br />

50<br />

0<br />

10 20 30 40 50<br />

% clay<br />

FS vs % clay<br />

Figure 10.12. Correlation between laboratory measured values <strong>of</strong> free swell <strong>an</strong>d<br />

clay fraction <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Figures (10.11) <strong>an</strong>d (10.12) illustrate a slight general increase <strong>of</strong> linear shrinkage <strong>an</strong>d free<br />

swell values with increasing clay content. The broad spread <strong>of</strong> results, however, points to the<br />

possible influence <strong>of</strong> other factors such as soil fabric <strong>an</strong>d/ or structure <strong>an</strong>d clay mineralogy on<br />

activity <strong>an</strong>d <strong>engineering</strong> behaviour <strong>of</strong> the soils.


200<br />

Chapter 11<br />

Conclusions<br />

This study has highlighted the presence <strong>of</strong> extensive reactive (potentially shrinking <strong>an</strong>d<br />

swelling) <strong>an</strong>d/ or exp<strong>an</strong>sive black cotton soils across the project area. Swelling capabilities<br />

(free swell values) <strong>of</strong> 100-170% (me<strong>an</strong>: 130%) <strong>an</strong>d swelling pressures <strong>of</strong> 49-104 kPa (me<strong>an</strong>:<br />

77 kPa) are exhibited by the soils.<br />

The study has also highlighted the influence <strong>of</strong> environmental factors such as rainfall, solar<br />

radiation, vegetation cover <strong>an</strong>d drainage on the nature <strong>an</strong>d behaviour <strong>of</strong> the soils in situ. Some<br />

sites in areas <strong>of</strong> black <strong>clays</strong> exhibited only a fraction <strong>of</strong> <strong>an</strong>ticipated shrinkage phenomena due<br />

to limited ch<strong>an</strong>ge in soil moisture conditions; whereas at other sites, extremely strong<br />

shrinkage cracking up to subst<strong>an</strong>tial depths <strong>an</strong>d in excess <strong>of</strong> that <strong>an</strong>ticipated based on soil<br />

index properties, has been observed <strong>an</strong>d recorded.<br />

A r<strong>an</strong>ge <strong>of</strong> soil index <strong>an</strong>d <strong>engineering</strong> properties as well as chemical/ mineralogical<br />

composition consistent with the <strong>geological</strong> diversity <strong>of</strong> the project area <strong>an</strong>d its surroundings,<br />

has been obtained <strong>an</strong>d recorded.<br />

The variability <strong>of</strong> index properties <strong>an</strong>d chemical/ mineralogical composition <strong>of</strong> black cotton<br />

soils is generally uniform with depth <strong>an</strong>d across the study area, implying occurrence <strong>of</strong><br />

generally homogeneous black <strong>clays</strong>.<br />

Most soil index properties could be characterised <strong>an</strong>d/ or estimated with a high degree <strong>of</strong><br />

approximation through correlation with results <strong>of</strong> Atterberg limits <strong>an</strong>d other index tests<br />

performed on disturbed <strong>an</strong>d fractioned samples.<br />

Characterisation <strong>of</strong> <strong>engineering</strong> behaviour <strong>of</strong> soils (shear strength, consolidation-settlement,<br />

swelling pressure & percentage swelling) is better accomplished by direct measurements on<br />

undisturbed core samples rather th<strong>an</strong> through correlation with results <strong>of</strong> indirect index tests<br />

performed on disturbed <strong>an</strong>d part samples which harbour destroyed soil fabric <strong>an</strong>d structure.<br />

The field <strong>an</strong>d laboratory investigations <strong>an</strong>d studies <strong>of</strong> soils <strong>of</strong> the project area have provided a<br />

vast data base on soil index <strong>an</strong>d <strong>engineering</strong> properties as well as chemical/ mineralogical<br />

composition. This data will provide the basis for on-going <strong>an</strong>d future research on the<br />

<strong>characterisation</strong> <strong>of</strong> exp<strong>an</strong>sive <strong>an</strong>d reactive black cotton soils, especially as regards their<br />

negative environmental, <strong>engineering</strong> <strong>geological</strong>/ civil <strong>engineering</strong> <strong>an</strong>d social-economic<br />

implications.<br />

There has been a progressive depletion in the proportionate amounts <strong>of</strong> soluble bases (MgO,<br />

CaO, Na2O, K2O) <strong>an</strong>d a corresponding increase in the proportions <strong>of</strong> iron <strong>an</strong>d aluminium<br />

compounds (Fe2O3 <strong>an</strong>d Al2O3) contained in the red soils with time. This is evidenced by<br />

comparing results <strong>of</strong> chemical <strong>an</strong>alyses for the red soils obtained in this study with previous<br />

chemical <strong>an</strong>alyses results for the same soils (Dumbleton, 1967; Sherwood, 1967). The<br />

depletion is most probably attributed to leaching <strong>of</strong> the soluble bases from the red soils,<br />

facilitated by the relatively good drainage conditions <strong>of</strong> the soils <strong>an</strong>d high rainfall received<br />

(760-1000 mm per year), thereby leaving the soils enriched in iron oxide <strong>an</strong>d haematite, as<br />

well as aluminium in the form <strong>of</strong> clay minerals.


201<br />

There has also been a decrease in the proportions <strong>of</strong> clay fraction as well as values <strong>of</strong><br />

Atterberg limits <strong>an</strong>d other index properties in the red soils with time. This has been evidenced<br />

by comparing results <strong>of</strong> soil classification tests obtained for the soils in this study with similar<br />

<strong>an</strong>d earlier results obtained for the same soils (Sherwood, 1967). This decrease could be<br />

explained by the increasing presence <strong>of</strong> free iron oxide in the soils, <strong>an</strong>d which has the effect<br />

<strong>of</strong> cementing the clay particles together, thereby causing a progressive decrease in both the<br />

proportions <strong>of</strong> clay as well as values <strong>of</strong> Atterberg limits <strong>of</strong> the red soils with time.<br />

Results <strong>of</strong> X-ray diffraction (XRD) <strong>an</strong>alyses show the mineralogical composition <strong>of</strong> red soils<br />

to be mainly kaolinite (80-81%) <strong>an</strong>d haematie (15-16%). Accessories <strong>of</strong> quartz (1-3%),<br />

tit<strong>an</strong>ium, i.e.ilmenite <strong>an</strong>d/ or rutile (~ 2%) <strong>an</strong>d m<strong>an</strong>g<strong>an</strong>ese (~ 1%) also occur. The kaolinite is<br />

depleted in Al2O3 <strong>an</strong>d this is probably caused by low PH conditions imparted to the soils by<br />

weak carbonic acids formed by processes <strong>of</strong> org<strong>an</strong>ic matter decomposition.<br />

The red soils encountered in this study are polygenetic; <strong>an</strong>d have been derived, developed <strong>an</strong>d<br />

formed under humid conditions, high temperatures <strong>an</strong>d presence <strong>of</strong> carbon dioxide by<br />

weathering <strong>an</strong>d alteration <strong>of</strong> Nairobi trachytes, volc<strong>an</strong>ic tuff <strong>an</strong>d ash. This suggestion is<br />

supported by comparing results <strong>of</strong> chemical <strong>an</strong>alyses obtained for the red soils in this study<br />

with similar results previously obtained for the underlying volc<strong>an</strong>ic materials; <strong>an</strong>d both <strong>of</strong><br />

which generally fall into close agreement. A limited contribution by eroded <strong>an</strong>d watertr<strong>an</strong>sported<br />

detrital materials is confirmed by results <strong>of</strong> sc<strong>an</strong>ning electron microscope which<br />

reveal inclusions <strong>of</strong> sub-rounded to rounded quartz grains. Leaching effects in the red soils<br />

<strong>an</strong>d consequent removal <strong>of</strong> silica <strong>an</strong>d soluble bases have caused the soils to exhibit lower<br />

contents <strong>of</strong> SiO2, MgO, CaO, Na2O, K2O <strong>an</strong>d correspondingly higher contents <strong>of</strong> Fe2O3 <strong>an</strong>d<br />

Al2O3 relative to the volc<strong>an</strong>ic materials. This is evidenced by <strong>an</strong>alysis results <strong>of</strong> sc<strong>an</strong>ning<br />

electron microscope for the red soils which show Fe-rich crusts as part <strong>of</strong> the matrix while<br />

quartz <strong>an</strong>d feldspars are in limited availability <strong>an</strong>d/or virtually absent.<br />

On the other h<strong>an</strong>d, results <strong>of</strong> X-ray diffraction <strong>an</strong>d sc<strong>an</strong>ning electron microscope <strong>an</strong>alyses<br />

show the clay minerals composition <strong>of</strong> black <strong>clays</strong> to be mainly smectites <strong>an</strong>d/ or<br />

montmorillonite (over 90%) <strong>an</strong>d some kaolinite (less th<strong>an</strong> 10%). Occurrence <strong>of</strong> illite is<br />

occasional <strong>an</strong>d only in trace form. Quartz (4-9%) <strong>an</strong>d accessories <strong>of</strong> K-feldspars (s<strong>an</strong>idine,<br />

orthoclase, microcline), haematite <strong>an</strong>d carbonates (calcite, dolomite, <strong>an</strong>kerite, siderite) also<br />

occur.<br />

Results <strong>of</strong> chemical <strong>an</strong>alyses together with current field observations <strong>an</strong>d evidences related to<br />

the nature <strong>of</strong> occurrence <strong>of</strong> black <strong>clays</strong> show that these soils are not wholly developed in situ;<br />

<strong>an</strong>d are not derived from the underlying practically impermeable Nairobi <strong>an</strong>d Kapiti<br />

phonolites through weathering <strong>an</strong>d alteration. Rather, the black <strong>clays</strong> have been derived from<br />

the gradual conversion <strong>of</strong> volc<strong>an</strong>ic ash as well as colluvium, alluvium <strong>an</strong>d s<strong>of</strong>t Pleistocene<br />

materials, all previously occurring as lacustrine deposits in a basin-like lake <strong>an</strong>d swampy<br />

environment during a pluvial period. Further confirmation is provided by results <strong>of</strong> sc<strong>an</strong>ning<br />

electron microscope which show presence <strong>of</strong> detrital components <strong>of</strong> biotite <strong>an</strong>d sub-rounded<br />

to rounded quartz grains as well as heavy minerals <strong>of</strong> Ti <strong>an</strong>d Zn in the black <strong>clays</strong>, a result <strong>of</strong><br />

erosion processes <strong>an</strong>d water tr<strong>an</strong>sport <strong>of</strong> soil/rock materials from surrounding areas. The black<br />

<strong>clays</strong> generally exhibit higher concentrations <strong>of</strong> soluble bases <strong>an</strong>d silica th<strong>an</strong> the red soils; <strong>an</strong>d<br />

this confirms the suggestion that leached components <strong>of</strong> soluble bases <strong>an</strong>d silica derived from<br />

areas <strong>of</strong> red soils occurring on higher grounds serve to further contribute to the formation <strong>of</strong><br />

black <strong>clays</strong> in the plains under alkaline <strong>an</strong>d impeded drainage conditions.


202<br />

The black <strong>clays</strong> <strong>an</strong>d red soils generally exhibit low total carbon contents <strong>of</strong> less th<strong>an</strong> 2,5%.<br />

The black <strong>clays</strong> show 0,50-1,48% total carbon (me<strong>an</strong>: 1,17%), 0,22-1,19% org<strong>an</strong>ic carbon<br />

(me<strong>an</strong>: 0,93%) <strong>an</strong>d 0,04-0,74% inorg<strong>an</strong>ic carbon (me<strong>an</strong>: 0,24%); while the red soils exhibit<br />

0,45-2,47% total carbon (me<strong>an</strong>: 1,12%), 0,35-1,90 org<strong>an</strong>ic carbon (me<strong>an</strong>: 0,85%) <strong>an</strong>d 0,10-<br />

0,57% inorg<strong>an</strong>ic carbon (me<strong>an</strong>: 0,27%). The black <strong>clays</strong> also show a general decrease <strong>of</strong> total<br />

carbon <strong>an</strong>d org<strong>an</strong>ic carbon, <strong>an</strong>d therefore, org<strong>an</strong>ic matter content <strong>of</strong> the soils with depth; the<br />

variation being best described <strong>an</strong>d represented by a polynomial (R = 0,56) <strong>an</strong>d exponential (R<br />

= 0,85) relationship, respectively. The inorg<strong>an</strong>ic carbon component generally increases with<br />

depth through in situ conversion <strong>of</strong> underlying volc<strong>an</strong>ic tuffs to secondary limestone; <strong>an</strong>d this<br />

variation tends to fit a polynomial distribution (R = 0,63). On the other h<strong>an</strong>d, the red soils<br />

show decreasing total carbon as well as org<strong>an</strong>ic <strong>an</strong>d inorg<strong>an</strong>ic carbon contents with depth, the<br />

three variations closely approximating a potential relationship with fairly strong correlations<br />

<strong>of</strong> R = 0,94. The decrease <strong>of</strong> the org<strong>an</strong>ic carbon is in reflection <strong>of</strong> decreasing org<strong>an</strong>ic matter<br />

content <strong>of</strong> the soils with depth; while decreasing inorg<strong>an</strong>ic carbon could be attributed to low<br />

pH conditions <strong>an</strong>d leaching <strong>of</strong> carbonates <strong>an</strong>d other soluble components <strong>of</strong> Mg, Ca, K <strong>an</strong>d Na<br />

from the soils during rainfall.<br />

Very s<strong>of</strong>t <strong>an</strong>d sensitive <strong>clays</strong> were also encountered in the present area. They occur in terms<br />

<strong>of</strong> black <strong>clays</strong> in peaty/ swampy environs <strong>an</strong>d/ or those <strong>of</strong> impended drainage; as well as red<br />

soils in densely forested <strong>an</strong>d uncultivated zones. The soils generally exhibit a very s<strong>of</strong>t to firm<br />

consistency, with v<strong>an</strong>e shear strength values <strong>of</strong> less th<strong>an</strong> 100 kPa. The undrained shear<br />

strength r<strong>an</strong>ges from 30,94 – 41,50 kPa for black <strong>clays</strong>; <strong>an</strong>d 10,92 – 87,36 kPa for red soils,<br />

<strong>an</strong>d generally increases with increased soil depths. The variation <strong>of</strong> v<strong>an</strong>e shear strength with<br />

soil depth closely fits a logarithmic relationship with a strong correlation (R = 0,94 for black<br />

<strong>clays</strong>; <strong>an</strong>d R = 0,97 – 1,0 for red soils). The variation <strong>of</strong> v<strong>an</strong>e shear strength with natural<br />

moisture content <strong>of</strong> soils is best described by a polynomial relationship with a very strong<br />

correlation (R = 1.0 for black <strong>clays</strong> <strong>an</strong>d R = 0,91 - 0,99 for red soils). Variation <strong>of</strong> v<strong>an</strong>e shear<br />

strength could also be predicted from bulk densities <strong>of</strong> soils through a polynomial relationship<br />

with a very strong correlation (R = 0,97 for black <strong>clays</strong>; R = 0,99-1,0 for red soils).<br />

A soils classification based on index properties <strong>an</strong>d grain sizes places the black <strong>clays</strong> into very<br />

high to extremely high plasticity <strong>clays</strong>, silty <strong>clays</strong> <strong>an</strong>d/ or silty <strong>clays</strong> with s<strong>an</strong>d, as well as<br />

clayey silts; all <strong>of</strong> which harbour medium to very high levels <strong>of</strong> activity (normal to highly<br />

active) so that they exhibit high swelling capabilities <strong>an</strong>d potential exp<strong>an</strong>siveness on wetting<br />

from a dry condition. This <strong>engineering</strong> behaviour has been explained by the high content <strong>of</strong><br />

the more exp<strong>an</strong>sive clay minerals, i.e. smectites (90% <strong>an</strong>d over) in the clay fraction. On the<br />

other h<strong>an</strong>d, the red soils are medium to high plasticity clayey silts <strong>an</strong>d silty <strong>clays</strong> with s<strong>an</strong>d;<br />

having at most only medium activity levels (normal active) <strong>an</strong>d therefore exhibiting generally<br />

low swelling capabilities <strong>an</strong>d potential exp<strong>an</strong>siveness when allowed free access to water.<br />

These characteristics have been attributed to the high content (80% <strong>an</strong>d over) <strong>of</strong> the rather less<br />

exp<strong>an</strong>sive clay mineral, kaolinite, in the clay fraction <strong>of</strong> the soils.<br />

The distribution <strong>an</strong>d variation <strong>of</strong> index properties <strong>of</strong> black <strong>clays</strong> across the study area at<br />

depths <strong>of</strong> less th<strong>an</strong> 0,50m <strong>an</strong>d those at 0,50m <strong>an</strong>d greater, were found to be generally similar,<br />

comparable <strong>an</strong>d in good agreement, implying a generally homogeneous <strong>engineering</strong> character<br />

<strong>of</strong> the soils. Index properties <strong>of</strong> red soils point to a comparatively more stable <strong>an</strong>d reliable<br />

<strong>engineering</strong> character <strong>of</strong> the soils.<br />

A new relationship has been derived, developed <strong>an</strong>d established in this study for the<br />

estimation <strong>of</strong> the plasticity index (PI) <strong>of</strong> black <strong>clays</strong> from laboratory determined linear<br />

shrinkage (LS) <strong>of</strong> the soils, i.e.


203<br />

PI = 1,88*LS<br />

A comparison <strong>of</strong> plasticity index values calculated using the above relationship (PILS) with<br />

those actually measured in the laboratory (PImeasured) has shown <strong>an</strong>d confirmed a strong<br />

agreement in the two sets <strong>of</strong> data in the form <strong>of</strong><br />

PImeasured = 1,005PILS<br />

The plasticity index <strong>of</strong> red soils could also be estimated from laboratory measured linear<br />

shrinkage using a new relationship developed for the soils, i.e.<br />

PI = 1,84LS<br />

The so calculated <strong>an</strong>d/ or estimated plasticity indices have also been found to be practically<br />

identical to actual laboratory measured values, i.e.<br />

PImeasured = 1,00PILS<br />

Another new relationship has been developed for the estimation <strong>of</strong> plasticity index (PI) <strong>of</strong><br />

black <strong>an</strong>d red soils from laboratory measured values <strong>of</strong> liquid limit (LL) with a high degree <strong>of</strong><br />

approximation, i.e.<br />

PI = 0,79(LL-25)<br />

The so calculated <strong>an</strong>d laboratory measured plasticity indices have been found to be in near<br />

perfect agreement, i.e.<br />

PImeasured = 0,99PILL<br />

Linear shrinkage values could also be estimated by calculation from laboratory determined<br />

liquid limits <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils using the newly developed relationship below, i.e.<br />

LS = 0,39(LL-21)<br />

The calculated <strong>an</strong>d laboratory measured linear shrinkage values have been found to be in very<br />

strong agreement, i.e.<br />

LSmeasured = 0,997LSLL<br />

The swelling capability <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils, in terms <strong>of</strong> free swell (FS), could be<br />

estimated with a high degree <strong>of</strong> approximation by calculation from laboratory determined<br />

Atterberg limits (LL, PI) <strong>an</strong>d linear shrinkage (LS). The new derived relationships take the<br />

form <strong>of</strong><br />

FS = 3,17(LL-45),<br />

FS = 3,79(PI-13) <strong>an</strong>d<br />

FS = 7,29(LS-7)


204<br />

The reliability <strong>of</strong> the estimations has been demonstrated by the close relationship <strong>an</strong>d strong<br />

agreement between laboratory measured <strong>an</strong>d so calculated free swell values, i.e.<br />

FSmeasured = 1,01FSLL<br />

FSmeasured = 0,99FSPI<br />

FSmeasured = 0,98FSLS<br />

Correlations between clay fraction/ content <strong>an</strong>d other index properties such as linear<br />

shrinkage <strong>an</strong>d free swell were found to be weak <strong>an</strong>d poor, most probably due to the effects <strong>of</strong><br />

soil fabric/ structure, clay mineralogy <strong>an</strong>d other factors in controlling the activity <strong>an</strong>d<br />

physical/ <strong>engineering</strong> behaviour <strong>of</strong> the soils.<br />

Laboratory shear strength investigations showed the black <strong>clays</strong> to be limited in their strength;<br />

<strong>an</strong>d are characterised by relatively lower values <strong>of</strong> <strong>an</strong>gles <strong>of</strong> shear resist<strong>an</strong>ce (φ`) <strong>of</strong> between<br />

11° <strong>an</strong>d 30°, giving <strong>an</strong> average value <strong>of</strong> 18°. The cohesive nature <strong>of</strong> the <strong>clays</strong> is reflected in<br />

their large cohesion (c´) values <strong>of</strong> 12 – 48 kN/m² (me<strong>an</strong> value: 35 kN/m²). On the other h<strong>an</strong>d,<br />

the red soils were found to be relatively more stable, with <strong>an</strong>gles <strong>of</strong> shear resist<strong>an</strong>ce <strong>of</strong> 28° to<br />

29°. However, the cohesive character <strong>of</strong> these soils is negligible due to their generally loose<br />

<strong>an</strong>d friable nature.<br />

Correlations between shear <strong>an</strong>gles determined on undisturbed samples <strong>an</strong>d results <strong>of</strong> index<br />

tests <strong>of</strong> black <strong>clays</strong> were only <strong>of</strong> poor to fair or moderate strength. A lot <strong>of</strong> uncertainty would<br />

therefore be encountered in attempting to assess <strong>an</strong>d characterise shear strength <strong>of</strong> the <strong>clays</strong><br />

on the basis <strong>of</strong> results <strong>of</strong> index properties obtained by tests carried out on disturbed <strong>an</strong>d<br />

fractioned samples.<br />

Results <strong>of</strong> laboratory oedometer consolidation tests <strong>an</strong>d compression coefficients derived<br />

there<strong>of</strong> show the red soils to be medium to very highly compressible (mv =0,16-2,94 m²/MN )<br />

when externally loaded in the r<strong>an</strong>ge <strong>of</strong> 25-800kPa; <strong>an</strong>d this is a probable result <strong>of</strong> their loose<br />

<strong>an</strong>d friable nature. The soils also exhibit medium to high rates <strong>of</strong> consolidation-settlement (cv<br />

= 1,18-12,14 m²/year), a result <strong>of</strong> the relatively high porosity (voids ratio e = 0,59-1,42) <strong>an</strong>d<br />

permeability (K = 7,57E-11 to 7,33E-9) which facilitate faster drainage <strong>an</strong>d rapid pore water<br />

pressure dissipation on external loading. In practice therefore, these soils would be expected<br />

to undergo rapid consolidation – settlements, especially during the construction stage, without<br />

posing long-term instability problems to constructed structures. The soils fall in the class <strong>of</strong><br />

normally consolidated to slightly overconsolidated <strong>clays</strong>.<br />

The black <strong>clays</strong> would exhibit slightly lower compressibility, i.e. medium to high (mv = 0,03-<br />

1,62 m²/MN) when externally loaded over a 25 –800kPa r<strong>an</strong>ge, as a result <strong>of</strong> their generally<br />

dense, compact <strong>an</strong>d cohesive nature. The <strong>clays</strong> would also exhibit low rates <strong>of</strong> consolidationsettlement<br />

(cv = 0,05-5,08 m²/year) due to their relatively low porosity (e = 0,23-1,06) <strong>an</strong>d<br />

permeability (K = 1,67E-12 to 1,20E-9) that tend to limit drainage <strong>an</strong>d dissipation <strong>of</strong> porewater<br />

pressures. Settlement <strong>an</strong>d instability <strong>of</strong> structures located on these <strong>clays</strong> would therefore<br />

be expected to persist beyond the construction stage. The <strong>clays</strong> are classified as normally<br />

consolidated.<br />

The following relationships have been derived in this study for the approximate estimation <strong>of</strong><br />

compressibility indices, cc, <strong>of</strong> the soils using laboratory determined liquid limits, LL, i.e.


205<br />

cc = 0,0099(122-LL)<br />

for black <strong>clays</strong>, <strong>an</strong>d<br />

cc = 0,0016(308-LL)<br />

for both black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Correlation between so calculated compression indices <strong>an</strong>d laboratory measured compression<br />

indices were found to be strong (R = 0,85) for the black <strong>clays</strong> alone; <strong>an</strong>d moderate (R = 0,58)<br />

for combined black <strong>clays</strong> <strong>an</strong>d red soils.<br />

Correlations between swell indices, Cs, <strong>an</strong>d index properties (Atterberg limits, free swell) <strong>of</strong><br />

black <strong>clays</strong> <strong>an</strong>d red soils were found to be generally poor. The black <strong>clays</strong> exhibit signific<strong>an</strong>t<br />

swelling pressures <strong>of</strong> 49-104 kPa. However, strengths <strong>of</strong> correlation between swelling<br />

pressures <strong>an</strong>d index properties were found to be generally weak. The above findings serve to<br />

highlight uncertainities that may be encountered in attempting to assess <strong>an</strong>d predict exp<strong>an</strong>sion<br />

tendencies (on unloading in situ) as well as swelling pressures imposed by clay soils on their<br />

saturation, based on results <strong>of</strong> laboratory index tests carried out on disturbed <strong>an</strong>d fractioned<br />

soils.<br />

Swelling test results have facilitated derivation <strong>of</strong> two methods <strong>an</strong>d/ or relationships that serve<br />

to summarise the swelling characteristics <strong>of</strong> black <strong>clays</strong>. Both relationships are logarithmic in<br />

representing the variation <strong>of</strong> percentage swelling (S%) with extent <strong>of</strong> external loading, P%<br />

(i.e., load decrements, P kPa, expressed as a percentage <strong>of</strong> swelling pressure, SP kPa). The<br />

first relationship involves deriving percentage swelling by expressing amount <strong>of</strong> swelling (S<br />

mm) as a fraction <strong>of</strong> the ultimate swelling (Smax mm) which occurs under zero external<br />

loading conditions. This relationship exhibits a very strong correlation (R = 0,995), i.e.<br />

P = -22,3Ln(S) + 100,80, where<br />

S (%) = percentage swelling (S mm/Smax mm)<br />

P (%) = imposed loads (P kPa) expressed as a percentage <strong>of</strong> swelling pressure (SP kPa)<br />

Based on this first relationship, therefore, it has been shown that black <strong>clays</strong> would undergo<br />

high percentage swelling (S >75%) when externally loaded to less th<strong>an</strong> 5 kPa; <strong>an</strong>d low<br />

percentage swelling for loads <strong>of</strong> over 80 kPa. As a result, external loads <strong>of</strong> well over 100 kPa<br />

would ensure minimal potential destabilisation <strong>of</strong> constructed light <strong>engineering</strong> structures<br />

from swelling effects <strong>of</strong> the <strong>clays</strong> on wetting. For known swelling pressure (SP) <strong>an</strong>d ultimate<br />

amount <strong>of</strong> swelling (Smax) under zero loading, the relationship could be possibly used as a<br />

guide to estimate external loading <strong>of</strong> light <strong>engineering</strong> structures necessary to give a desired<br />

<strong>an</strong>d/ or permitted percentage swelling, as may be required by design computations. However,<br />

the relationship c<strong>an</strong>not be reliably used for estimation <strong>of</strong> percentage swelling from known<br />

values <strong>of</strong> external loads, since the former is usually underestimated by <strong>an</strong> average amount <strong>of</strong><br />

S = 8%.<br />

The second relationship has percentage swelling derived by relating amount <strong>of</strong> swelling, S<br />

mm, to initial test specimen thickness, Ho mm. It is also closely logarithmic with a similarly<br />

very strong correlation (R = 0,98), i.e.<br />

P = -15,86Ln(S) + 29,84


206<br />

where P (%) = (P kPa/SP kPa) * 100<br />

S (%) = (S mm/Ho mm) * 100<br />

Unlike the first relationship, this last relationship could be used for estimation <strong>of</strong> both external<br />

loading conditions necessary to meet <strong>an</strong> allowable percentage swelling <strong>an</strong>d vice versa, as long<br />

as the swelling pressure <strong>of</strong> the <strong>clays</strong> is known. It would therefore be a preferred relationship in<br />

estimating <strong>an</strong>d/ or characterising swelling characteristics <strong>of</strong> <strong>clays</strong>.


207<br />

Chapter 12<br />

Recommendations<br />

Analysis <strong>of</strong> short term stability problems during <strong>an</strong>d/ or immediately after construction<br />

involve use <strong>of</strong> the undrained shear strength or apparent cohesion, cu, in the relev<strong>an</strong>t<br />

computations (Terzaghi <strong>an</strong>d Peck, 1967). The corresponding <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce, φ, is<br />

especially useful in the derivation <strong>of</strong> the necessary earth pressure coefficients <strong>an</strong>d/ or bearing<br />

capacity coefficients. It would be more helpful therefore, that future works involving soils <strong>of</strong><br />

the area include laboratory triaxial tests in their studies with the purpose <strong>of</strong> providing the<br />

necessary undrained shear strength parameters, cu, φ. The triaxial testing procedure is<br />

generally more satisfactory in measuring the shear strength <strong>of</strong> clay soils in terms <strong>of</strong> total<br />

stresses, through quick undrained shear tests (Head, 1988). Short-term stability <strong>an</strong>alyses may<br />

include computations related to bearing capacity <strong>of</strong> footings <strong>an</strong>d foundations for structures on<br />

saturated homogeneous <strong>clays</strong>, earth pressure on retaining walls, earth pressure against bracing<br />

in temporary excavations, safeguard against heave <strong>of</strong> the bottom <strong>of</strong> temporary open<br />

excavations in <strong>clays</strong>, stability <strong>of</strong> side-slopes <strong>of</strong> cuttings, as well as short term stability <strong>of</strong><br />

emb<strong>an</strong>kments <strong>an</strong>d earth dams.<br />

The black <strong>clays</strong> in some parts <strong>of</strong> the study area were found to contain gravelly lateritic<br />

materials <strong>of</strong> up to 40 mm sizes. Investigation <strong>an</strong>d determination <strong>of</strong> shear strength<br />

characteristics <strong>of</strong> such materials by triaxial testing would be impracticable; while use <strong>of</strong> the<br />

st<strong>an</strong>dard shearbox apparatus as employed in this study, would only produce unreliable results.<br />

Future investigations <strong>an</strong>d studies regarding shear strength <strong>of</strong> these signific<strong>an</strong>tly coarse<br />

gravelly <strong>clays</strong> should therefore employ the large shearbox apparatus such as the one described<br />

by Head (1988). This apparatus could provide more reliable shear strength parameters by<br />

executing slow drained shear tests on large representative <strong>an</strong>d undisturbed block samples <strong>of</strong><br />

the <strong>clays</strong> measuring 305 mm square <strong>an</strong>d 150 mm thick. The results <strong>of</strong> shear strength<br />

parameters so obtained would be useful at the design stages <strong>of</strong> a number <strong>of</strong> structures such as<br />

emb<strong>an</strong>kments <strong>an</strong>d earth dams which usually incorporate gravel fill material (Pike, 1973;<br />

Pike, Acott <strong>an</strong>d Leech, 1977). The results would also serve as a basis for discriminating,<br />

subdividing <strong>an</strong>d classifying gravelly soil materials to suit different construction needs <strong>an</strong>d<br />

works (Bishop, 1948; Pike, 1973).<br />

The red soils were found in this study to be generally more permeable <strong>an</strong>d relatively free<br />

draining. They would therefore exhibit rapid drainage <strong>an</strong>d settlement when externally loaded.<br />

The resulting compression/ settlement curves tend to deviate from conventional shapes<br />

derived from one-dimensional theory <strong>of</strong> consolidation so that determination <strong>of</strong> relev<strong>an</strong>t<br />

consolidation parameters would also be difficult. Use <strong>of</strong> a Rowe consolidation cell<br />

(Rowe,1966), capable <strong>of</strong> accommodating larger soil specimens would be recommended for<br />

oedometer tests on the clayey silt <strong>an</strong>d silty varieties <strong>of</strong> red soils, so as to obtain a more<br />

definite value <strong>of</strong> coefficient <strong>of</strong> consolidation, cv, for use in the more reliable estimation <strong>of</strong><br />

rates <strong>of</strong> settlement under various loading conditions. Alternatively, cv could be derived<br />

empirically from the relation<br />

cv = K/(mv * 0,31*E-9) m²/year<br />

where K (m/s) = coefficient <strong>of</strong> permeability as measured in situ in the field<br />

mv (MN/m²) = coefficient <strong>of</strong> volume compressibility derived from laboratory<br />

consolidaton test using the Rowe cell


208<br />

In addition, more reasonable estimates <strong>of</strong> both the amount <strong>an</strong>d rate <strong>of</strong> settlement <strong>of</strong> soils <strong>of</strong><br />

the project area could be achieved by taking horizontal drainage into account during<br />

laboratory consolidation tests. This could be accomplished by performing consolidation tests<br />

on larger specimens under hydraulic loading using a larger cell, such as the Rowe<br />

consolidation cell, <strong>an</strong>d designed for provision <strong>of</strong> horizontal drainage (Rowe, 1966). In the<br />

normal oedometer consolidation tests, however, horizontal drainage could be taken into<br />

account by either fitting a pervious lining inside the consolidation ring <strong>an</strong>d sealing the<br />

specimen ends; or by trimming the specimen in a vertical pl<strong>an</strong>e (Head, 1988).<br />

A provision should be made to determine in situ <strong>an</strong>d/ or monitor in the field consolidation<br />

settlements <strong>an</strong>d accomp<strong>an</strong>ying increase in effective strength which occur due to dewatering<br />

<strong>an</strong>d/ or partial drying <strong>of</strong> s<strong>of</strong>t soils (in swamp <strong>an</strong>d peaty environments) caused by a fall in the<br />

groundwater table; <strong>an</strong>d this especially in the tr<strong>an</strong>sition period from wet to dry months, <strong>an</strong>d<br />

vice versa. This would serve to assess <strong>an</strong>d predict possible destructive implications onto<br />

constructed <strong>an</strong>d/ or projected <strong>engineering</strong> structures. According to Head (1988), a fall <strong>of</strong> 1 m<br />

in the groundwater table would increase the effective stress in the whole clay deposit beneath<br />

the water table by about 10 kPa, the actual amount <strong>of</strong> consolidation depending on the ch<strong>an</strong>ge<br />

in effective stress in the soil.<br />

It has been established in this study that the plasticity index (PI) <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils<br />

could be approximately estimated from laboratory measured results <strong>of</strong> linear shrinkage (LS)<br />

<strong>an</strong>d/ or liquid limit (LL) using newly developed relationships, i.e.<br />

PI = 1,88*LS (for black <strong>clays</strong>),<br />

PI = 1,84LS (for red soils) <strong>an</strong>d<br />

PI = 0,79(LL-25) for both black <strong>clays</strong> <strong>an</strong>d red soils.<br />

In addition, the swelling characteristics <strong>of</strong> black <strong>clays</strong> could be approximately expressed in<br />

terms <strong>of</strong> logarithmic relationships derived in the present study, i.e.<br />

P = -22,3Ln(S) + 100,80, where<br />

S (%) = percentage swelling with respect to ultimate swelling (Smax mm) under zero<br />

loading<br />

P (%) = imposed loads (P kPa) expressed as a percentage <strong>of</strong> swelling pressure (SP kPa)<br />

<strong>an</strong>d P = -15,86Ln(S) + 29,84, for<br />

S (%) = percentage swelling with respect to initial test specimen thickness (Ho mm)<br />

P (%) = imposed loads (P kPa) expressed as a percentage <strong>of</strong> swelling pressure (SP kPa)<br />

It would be necessary <strong>an</strong>d useful to extend the investigation with the purpose <strong>of</strong> finding out if<br />

the same relationships for estimating plasticity index <strong>an</strong>d swelling characteristics hold for<br />

similar <strong>tropical</strong> <strong>clays</strong> <strong>an</strong>d/ or soils in Africa, <strong>an</strong>d around the world in general.<br />

The following relationships have been derived in this study for the approximate estimation <strong>of</strong><br />

compressibility indices, cc, <strong>of</strong> the soils using laboratory determined liquid limits, LL, i.e.


209<br />

cc = 0,0099(122-LL)<br />

for black <strong>clays</strong>, <strong>an</strong>d<br />

cc = 0,0016(308-LL)<br />

for both black <strong>clays</strong> <strong>an</strong>d red soils, where<br />

cc = a dimensionless number, <strong>an</strong>d<br />

LL (%) = liquid limit<br />

Correlation between so calculated compression indices <strong>an</strong>d laboratory measured compression<br />

indices were found to be strong (R = 0,85) for the black <strong>clays</strong> alone; <strong>an</strong>d moderate (R = 0,58)<br />

for combined black <strong>clays</strong> <strong>an</strong>d red soils. However, these derivations were based on a limited<br />

amount <strong>of</strong> data on compression indices (n = 6 for black <strong>clays</strong>; n = 9 for black <strong>clays</strong> <strong>an</strong>d red<br />

soils) available in the current study. It would be useful that future works attempt the same<br />

correlations <strong>an</strong>d derivations based on a larger amount <strong>of</strong> data as regards laboratory results <strong>of</strong><br />

liquid limit tests performed on disturbed samples on one h<strong>an</strong>d, <strong>an</strong>d those <strong>of</strong> compression<br />

indices obtained from oedometer tests on undisturbed specimens on the other. Possible<br />

improvements in the strength <strong>of</strong> correlations <strong>an</strong>d assessment <strong>of</strong> compressibility characteristics<br />

<strong>of</strong> soils using index properties (Atterberg limits) would also be indicated. Investigations could<br />

also be extended to other types <strong>of</strong> soil world-wide, to find out if the same <strong>an</strong>d/ or similar<br />

relationships hold.<br />

All in all, it may be useful to initiate <strong>an</strong>d establish a field monitoring program for the soils <strong>of</strong><br />

the Nairobi area, aimed at monitoring <strong>of</strong> exp<strong>an</strong>sive/ reactive soil movements <strong>an</strong>d moisture<br />

ch<strong>an</strong>ges. The selected sites would involve soils <strong>of</strong> varying geographic <strong>an</strong>d lithological<br />

conditions. This would serve the purpose <strong>of</strong> providing a comprehensive picture <strong>of</strong> exp<strong>an</strong>sive<br />

<strong>an</strong>d/ or reactive soil behaviour, in terms <strong>of</strong> <strong>characterisation</strong> <strong>of</strong> site reactivity (shrink-swell<br />

potential) <strong>an</strong>d unsaturated moisture flow. The scheme my take the form <strong>of</strong> monitoring <strong>an</strong>d<br />

recording pr<strong>of</strong>iles <strong>of</strong> ground movements <strong>an</strong>d moisture ch<strong>an</strong>ges as well as providing general<br />

data on soils with time (e.g. on a monthly basis), over a period <strong>of</strong> some years. A further<br />

<strong>an</strong>alysis in terms <strong>of</strong> environmental factors would be necessary <strong>an</strong>d helpful.<br />

According to Del<strong>an</strong>ey & Allm<strong>an</strong> (1998), useful instrumentation for a field monitoring<br />

program may include use <strong>of</strong> surface <strong>an</strong>d sub-surface pegs (surface <strong>an</strong>d sub-surface levels),<br />

neutron probes (soil volumetric moisture content), thermocouples (soil temperature), gypsum<br />

blocks (matrix soil suction), piezometer (groundwater level, especially in alluvial soils),<br />

automatic weather stations (rainfall, humidity, temperature, solar radiation, wind velocity/<br />

direction) etc. Results may include data on design parameters for foundations on exp<strong>an</strong>sive<br />

<strong>an</strong>d/ or reactive soils (e.g. active depth, suction ch<strong>an</strong>ge, seasonal heave etc.), as well as<br />

detailed information on field behaviour to compare with models which may be developed in<br />

the laboratory to simulate <strong>an</strong>d characterise possible field behaviour <strong>of</strong> the soils. The results<br />

would also provide a basis for comparing <strong>an</strong>d assessing the different methods <strong>of</strong><br />

characterising site reactivity <strong>an</strong>d/ or shrink-swell behaviour.<br />

The first <strong>of</strong> the two methods developed in this study to estimate swelling characteristics <strong>of</strong><br />

black <strong>clays</strong> [i.e. Eq. (7.49): P = -22,3Ln(S) + 100,8] was shown to be limited in predicting<br />

percentage swelling that may be realised under selected structural loading conditions. It was<br />

also explained that the underestimation <strong>of</strong> percentage swelling values was most probably a<br />

result <strong>of</strong> some degree <strong>of</strong> plastic deformation (or consolidation <strong>of</strong> test specimens when initially


210<br />

loaded to their swelling pressure), <strong>an</strong>d which could not therefore be recovered on subsequent<br />

complete unloading. In order to improve on the estimation capability <strong>an</strong>d reliability <strong>of</strong> this<br />

method, it is recommended that the amount <strong>of</strong> ultimate swelling (Smax) under zero loading,<br />

<strong>an</strong>d on which calculation <strong>of</strong> percentage swelling is based, be determined on separate test<br />

specimens from those used in the swelling test. Cumulative swelling values (S mm) obtained<br />

during the swelling test would then be expressed as a percentage <strong>of</strong> this separately determined<br />

maximun swelling value, Smax (mm).<br />

Adoption <strong>of</strong> a modified triaxial compression cell (after Abebe, 2002) to investigate exp<strong>an</strong>sive<br />

<strong>clays</strong> could prove useful in providing more accurate information on swelling characteristics<br />

through simult<strong>an</strong>eous measurement <strong>of</strong> swelling pressure <strong>an</strong>d associated swelling strain<br />

(volume ch<strong>an</strong>ge).


211<br />

References<br />

Abebe, S. T., 2002. Foundation pits in saturated highly exp<strong>an</strong>sive soils. Ph.D. thesis, Dept. Civil<br />

Engineering, Univ. Kaiserslautern, Germ<strong>an</strong>y.<br />

Allum, J.A.E., 1966. Photogeology <strong>an</strong>d regional mapping. Pergamon Press Ltd.<br />

Americ<strong>an</strong> Society for Testing <strong>an</strong>d Materials, 1979. Test Designation D 3080. St<strong>an</strong>dard<br />

method for direct shear test <strong>of</strong> soils under consolidated drained conditions. ASTM,<br />

Philadelphia, USA.<br />

-----, Part II. Test Designation D2435-70. One-dimensional consolidation properties <strong>of</strong> soils.<br />

ASTM, Philadelphia.<br />

Anyumba, J. <strong>an</strong>d Ruigu, F., 1991. Site investigation report for Plot No. L.R. 209/6869,<br />

Nairobi. Hydrodata Consult.<br />

Atterberg, A., 1911. Über die physikalische Bodenuntersuchung und über die plastizität der<br />

Tone. Internationale Mitteilungen für Bodenkunde, Vol. 1.<br />

Attewell, P.B. <strong>an</strong>d Farmer, I.W., 1976. Principles <strong>of</strong> <strong>engineering</strong> geology. Chapm<strong>an</strong> <strong>an</strong>d Hall,<br />

London.<br />

Baker, B. H., 1954. Geology <strong>of</strong> the south Machakos area. Rep. geol. Surv. Kenya 27.<br />

-----, 1958. Geology <strong>of</strong> the Magadi area. Rep. geol. Surv. Kenya 42.<br />

-----, Mitchell, J. G. <strong>an</strong>d Williams, L. A. J., 1988. Stratigraphy, geochronology <strong>an</strong>d<br />

volc<strong>an</strong>o – tectonic evolution <strong>of</strong> the Kedong-Naivasha-Kin<strong>an</strong>gop region, Gregory Rift<br />

Valley, Kenya. Q. J. geol. Soc. London 145, 107-117.<br />

Barden, L., 1965. Consolidation <strong>of</strong> compacted <strong>an</strong>d unsaturated <strong>clays</strong>. Geotechnique Vol. 15, No. 3.<br />

Bell, D.H. <strong>an</strong>d Pettinga, J.R., 1985. Engineering geology <strong>an</strong>d subdivision pl<strong>an</strong>ning in New<br />

Zeal<strong>an</strong>d. Eng. Geol., 22: 45-59.<br />

BGR (Bundes<strong>an</strong>stalt für Geowissenschaften und Rohst<strong>of</strong>fe), 1996. Bodenkundliche<br />

Kartier<strong>an</strong>leitung. E. Schweizerbart`scheVerlagsbuchh<strong>an</strong>dlung, Stuttgart.<br />

Bishop, A. W., 1948. A large shearbox for testing s<strong>an</strong>ds <strong>an</strong>d gravels. Proc. 2 nd Int. Conf. Soil<br />

Mech. & Found. Eng., Rotterdam,Vol. 1.<br />

Bjerrum, L., 1974. Problems <strong>of</strong> soil mech<strong>an</strong>ics <strong>an</strong>d construction on s<strong>of</strong>t <strong>clays</strong>. Norges Geotekn.<br />

Inst. Publ. Oslo. No. 100.<br />

Blyth, F.G.H. <strong>an</strong>d de Fretais, M.H., 1974. A Geology for Engineers. Edward Arnold Publ. Co.<br />

Bowen, N. L, 1937. Recent high-temperature research on silicates <strong>an</strong>d its signific<strong>an</strong>ce in<br />

igneous petrology. Am. J. Sci. 33, 1-21.<br />

------ <strong>an</strong>d Ellestad, 1936. Nepheline contrasts. Am. Min. 21, 363.


212<br />

BS Code <strong>of</strong> Practice CP 2001, 1957. Site investigations. British St<strong>an</strong>dards Institution, London; <strong>an</strong>d<br />

document No. 76/ 11937, draft revision.<br />

BS 1377:1967. Determination <strong>of</strong> linear shrinkage characteristics <strong>of</strong> clay soils. British St<strong>an</strong>dards<br />

Institution, London.<br />

BS 1377:1975. Laboratory determination methods <strong>of</strong> <strong>engineering</strong> soil properties. British<br />

St<strong>an</strong>dards Institution, London.<br />

Casagr<strong>an</strong>de, A., 1958. Notes on the design <strong>of</strong> the liquid limit device. Geotechnique, Vol. 8, No. 2.<br />

------,1936. The determination <strong>of</strong> the pre-consolidation load <strong>an</strong>d its practical signific<strong>an</strong>ce.<br />

Proc. 1 st Int. Conf. Soil Mech.,Cambridge, Mass., Vol. 3.<br />

-----,1932. The structure <strong>of</strong> clay <strong>an</strong>d its import<strong>an</strong>ce in foundation Engineering. J.<br />

Boston Soc. Civ. Eng., Vol. 19.<br />

Cheeney, R. F., 1983. Statistical methods in geology for field <strong>an</strong>d laboratory decisions. George<br />

Allen & Unwin Publishers Ltd., London, UK.<br />

Cochr<strong>an</strong>, W. G., 1977. Sampling techniques, 5. New York: Wiley.<br />

Coduto, D.P., 1994. Foundation Design, Principles <strong>an</strong>d Practices. Prentice-Hall, Englewood Cliffs,<br />

NJ.<br />

Colombo, P., 1983. Elementi di Geotecnica. Z<strong>an</strong>ichelli Editore S.p.A, Bologna.<br />

Collie, E. L, 1912. Plateau <strong>of</strong> British East Africa. Bull. geol. Soc. Am. 23, 297-316.<br />

Colwell, R.N., 1983. M<strong>an</strong>ual <strong>of</strong> remote sensing. 2nd ed.<br />

Davis, E. H. <strong>an</strong>d Poulos, H., 1965. The <strong>an</strong>alysis <strong>of</strong> settlement under three-dimensional<br />

conditions. Symp. on S<strong>of</strong>t Groud Eng., Inst. Eng. Australia, Brisb<strong>an</strong>e.<br />

Day, Robert W., 2001. Soil Testing M<strong>an</strong>ual: procedures, classification data <strong>an</strong>d sampling<br />

practices. Mcgraw-Hill.<br />

Del<strong>an</strong>ey, M. D., Allm<strong>an</strong>, M. A., Smith, D. W. <strong>an</strong>d Slo<strong>an</strong>, S. W., 1998. The establishment <strong>an</strong>d<br />

monitoring <strong>of</strong> exp<strong>an</strong>sive soil field sites. Geotechnical Site Characterisation, Balkema, Rotterdam.<br />

Deutsches Institut für Normung (DIN) e.V., 1980. Flügelsondierung nach DIN 4096-FS 50.<br />

DIN Taschenbuch 113, 1998. Erkundung und Untersuchung des Baugrunds.<br />

Dumbleton, M. J., 1967. Origin <strong>an</strong>d mineralogy <strong>of</strong> Afric<strong>an</strong> red <strong>clays</strong> <strong>an</strong>d Keuper Marl. Q. J.<br />

Eng. Geol. 1, 39-45.<br />

Evernden, J. F., Curtiss, G. H. <strong>an</strong>d Miller, 1964. Potassium-argon dates <strong>of</strong> volc<strong>an</strong>ic rocks<br />

from Nairobi <strong>an</strong>d the Rift region, Kenya. Am. J. Sci. 262, 145-198.<br />

Fa.Point elektronic GmbH Halle/Saale, 1997. Hardware (REM PC Einsteckkarte) for Digital<br />

Photo-processing <strong>an</strong>d S<strong>of</strong>tware (WinDISS 2.0) for Digital Photography.<br />

Fairburn, W.A., 1963. Geology <strong>of</strong> the North Machakos - Thika Area. Rep. 59, Geol. Surv. Kenya.


213<br />

Galster, R.W., 1977. A system <strong>of</strong> <strong>engineering</strong> geology mapping symbols. Bull. Int. Assoc. Eng.<br />

Geol., 14: 39-47.<br />

Gethin - Jones, G.H., 1953. A soil survey <strong>of</strong> the C<strong>of</strong>fee Research Station, Ruiru. Rep. Soil Surv.<br />

Kenya (not numbered).<br />

Gevaerts, E.A.L.,1964. Hydrogeology <strong>of</strong> the Nairobi area. Tech. Rep. 1. Water Dev. Dept.,<br />

Nairobi, Kenya.<br />

Gibbs, H. J. <strong>an</strong>d Holtz, W. G., 1956. Engineering properties <strong>of</strong> exp<strong>an</strong>sive <strong>clays</strong>. Tr<strong>an</strong>s. Am. Soc.<br />

Civ. Eng., Vol. 121, No.1, Paper 2814.<br />

Glossop, R. <strong>an</strong>d Skempton, A. W., 1945. Particle size in silts <strong>an</strong>d s<strong>an</strong>ds. Proc. Inst. Civ. Eng.<br />

Paper No. 5492.<br />

Gregory, J.W., 1921. The Great Rift Valley <strong>an</strong>d Geology <strong>of</strong> East Africa. Seeley Service, London.<br />

Häkli, A., 1960. On high – temperature alkali feldspars <strong>of</strong> some volc<strong>an</strong>ic rocks <strong>of</strong> Kenya <strong>an</strong>d<br />

northern T<strong>an</strong>g<strong>an</strong>yika. Compte Rendu Soc. geol. Finl<strong>an</strong>d 32, 99-108.<br />

Head, K. H., 1988. M<strong>an</strong>ual <strong>of</strong> soil laboratory testing. Vol. 2: permeability, shear strength <strong>an</strong>d<br />

compressibility tests. Pentech Press Ltd., Plymouth, Devon.<br />

-------,1984. M<strong>an</strong>ual <strong>of</strong> soil laboratory testing. Vol. 1: soil classification <strong>an</strong>d compaction tests.<br />

Pentech Press, London.<br />

Hedberg, H., 1976. An international guide to stratigraphic classification, terminology <strong>an</strong>d<br />

usage. Lethia, 5: 297-323.<br />

Holtz, R. D. <strong>an</strong>d Kovacs, W. D., 1981. An Introduction to Geotechnical Engineering. Prentice-<br />

Hall, Englewood Cliffs, NJ.<br />

IAEG, 1981a. Recommended symbols for <strong>engineering</strong> <strong>geological</strong> mapping. Bull. Int. Assoc. Eng.<br />

Geol., 24: 227-234.<br />

-------, 1981b. Rock <strong>an</strong>d soil description for <strong>engineering</strong> <strong>geological</strong> mapping. Bull. Int. Assoc.<br />

Eng. Geol., 24: 235-274.<br />

-------, 1976. Engineering <strong>geological</strong> maps: a guide to their preparation. UNESCO, Paris, 79 pp.<br />

ISI (International Scientific Instruments), 1972. Sc<strong>an</strong>ning Electron Microscope Leiz, ISI<br />

Super 40/3A.<br />

ISRM, 1978a. Suggested method for the petrographic description <strong>of</strong> rocks. Int. J. Rock Mech.<br />

Min. Sci. Geomech. Abstr., 15: 41-45.<br />

-------, 1978b. Suggested methods for the qu<strong>an</strong>titative descriptions <strong>of</strong> discontinuities in rock<br />

masses. Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 15: 319-368.<br />

Jasmund, K. <strong>an</strong>d Lagaly, G., 1993. Tonminerale und Tone: Struktur, Eigenschaften,<br />

Anwendung und Einsatz in Industrie und Umwelt. Steinkopff Verlag, Darmsatdt.<br />

Johnson, Robert B. <strong>an</strong>d Degraff, Jerome V., 1988. Principles <strong>of</strong> Engineering Geology. John<br />

Wiley & Sons, Inc.<br />

Joubert, P., Unpublished. Building materials <strong>of</strong> Nairobi. Mem. geol. Surv. Kenya.


214<br />

Kaye, G. W. C. <strong>an</strong>d Laby, T. H., 1973. Tables <strong>of</strong> Physical <strong>an</strong>d Chemical Const<strong>an</strong>ts, 14 th<br />

edition. Longm<strong>an</strong>s, London.<br />

Kenya Meteorological Department, 2001. The climate <strong>of</strong> Nairobi city <strong>an</strong>d surrounding areas.<br />

Koch, G. S. <strong>an</strong>d Link, R. F., 1970. Statistical <strong>an</strong>alysis <strong>of</strong> <strong>geological</strong> data. Vol. 1. New York:<br />

Wiley.<br />

Krenkel, E, 1925. Geologie Afrikas, Vol. 1. Berlin: Borntaeger.<br />

Krumbein, W. C. <strong>an</strong>d Graybill, F. A., 1965. An introduction to statistical models in geology.<br />

New York: McGraw-Hill.<br />

Kunzli, E, 1901. Die petrographische ausbente der Schöllerschen expedition in Aequatorial –<br />

Östafrica (Masail<strong>an</strong>d).. Vjschr naturf. Ges. Zurich 46, 128-172.<br />

Lacroix, A., 1923. Mineralogie de Madagascar 3. Paris: Challamel.<br />

Lambe, T. W. <strong>an</strong>d Whitm<strong>an</strong>, R. V., 1969. Soil Mech<strong>an</strong>ics. Wiley, New York.<br />

-----, 1979. Soil Mech<strong>an</strong>ics. S.I. Version, Wiley, New York.<br />

Leonards, G. A. (ed.), 1962. Foundation Engineering, Chapter 2. McGraw-Hill, New York.<br />

Lewis, W. A. <strong>an</strong>d Ross, N. F., 1955. An investigation <strong>of</strong> the relationship between the shear<br />

strength <strong>of</strong> remoulded cohesive soil <strong>an</strong>d the soil moisture suction. Unpublished Research Note No.<br />

RN/2389/WAL.NFR. Road Research Laboratory, Crowthorne, Berks.<br />

Maharaj, R.J., 1995. Engineering - <strong>geological</strong> mapping <strong>of</strong> <strong>tropical</strong> soils <strong>of</strong> Jamaica. Eng. Geol.,<br />

40: 243-286.<br />

M<strong>an</strong>ga, S.R., 1988. Site investigation report for I.C.D.C Investment Building, Uhuru<br />

Highway, Nairobi.<br />

Maufe, H. B, 1908. Reports relating to the geology <strong>of</strong> the East Africa Protectorate. Colon. Rep.<br />

misc. Ser. 45.<br />

Matheson, F. J., 1966. Geology <strong>of</strong> the Kajiado area. Rep. geol. Surv. Kenya 70.<br />

McCall, G. J. H., 1962. Froth-flow lavas resembling ignimbrites in the East Afric<strong>an</strong> Rift<br />

Valley. Nature 194, 343-344.<br />

-----, 1964. Froth flows in Kenya. Geol. Rundsch. 54, 1148-1195.<br />

MacDonald, D. H. <strong>an</strong>d Skempton, A. W., 1955. A survey <strong>of</strong> comparisons between calculated<br />

<strong>an</strong>d observed settlements <strong>of</strong> structures on clay. Paper No. 19. Conf. on Correlation Between<br />

Calculated <strong>an</strong>d Observed Stresses <strong>an</strong>d Displacements in Building. Inst. Civ. Eng., London.<br />

Meyer, S. L., 1975. Data <strong>an</strong>alysis for scientists <strong>an</strong>d engineers, 167-9, 342-9. New York: Wiley.<br />

Mines <strong>an</strong>d Geological Department, Nairobi, 2001. Unpublished <strong>geological</strong> works covering<br />

Nairobi <strong>an</strong>d its environs.


215<br />

Mitchell, J.K., 1993. Fundamentals <strong>of</strong> soil behaviour. 2nd ed., Wiley, N.Y., 437 pp.<br />

-----, 1976. Fundamentals <strong>of</strong> soil behaviour. Wiley, New York.<br />

-----, Vivatrat, V. <strong>an</strong>d Lambe, T. W., 1977. Foundation perform<strong>an</strong>ce <strong>of</strong> the tower <strong>of</strong><br />

Pisa. Proc. ASCE. Geotech. Eng. Div., Vol. 103, No. GT3.<br />

Moore, D. M. <strong>an</strong>d Reynolds, R. C., 1989. X-ray Diffraction <strong>an</strong>d the Identification <strong>an</strong>d<br />

<strong>an</strong>alysis <strong>of</strong> Clay Minerals. Oxford Univ. Press, Inc., New York.<br />

Morg<strong>an</strong>, W. T. W, ed, 1967. Nairobi: city <strong>an</strong>d region. Oxford Univ. Press, London.<br />

Mountain, E. D., 1925. Potash oligoclase from Mt. Erebus, Antarctica <strong>an</strong>d <strong>an</strong>orthoclase from<br />

Mt. Kenya, East Africa. Min. Mag. 20.<br />

Mowlem Construction Co. Ltd., (1985). Site investigation report for proposed headquarters <strong>of</strong><br />

Ukulima Cooperative Savings & Credit Society, Nairobi.<br />

Mussett, A. E., Reilly, T. A. <strong>an</strong>d Raja, P. K. S., 1964. Palaeomagnetism in East Africa. Report<br />

on the geology <strong>an</strong>d geophysics <strong>of</strong> the East Afric<strong>an</strong> Rift system.<br />

--- -, 1965. Palaeomagnetism in East Africa. A progress report on the Tertiary volc<strong>an</strong>ics.<br />

Proc. E. Afr. Acad. 2, 27-35.<br />

NAVFAC, 1982. Design m<strong>an</strong>ual: soil mech<strong>an</strong>ics. US Dept. <strong>of</strong> Defence, NAVFAC DM-7.1, Dept. <strong>of</strong><br />

the Navy, Washington, D.C., 360 pp.<br />

Nairobi City Council, 1999. Statistics on growth <strong>of</strong> Nairobi city <strong>an</strong>d its population.<br />

Nelson, John D. <strong>an</strong>d Miller, Debora J., 1992. Exp<strong>an</strong>sive Soils: Problems <strong>an</strong>d Practice in<br />

Foundation <strong>an</strong>d Pavement Engineering. John Wiley & Sons, Inc.<br />

Nevill, D., 1961. A laboratory investigation <strong>of</strong> two red <strong>clays</strong> from Kenya. Geotechnique 11, 302-<br />

318.<br />

NLB (Niedersächsischen L<strong>an</strong>desamt für Bodenforschung) und BGR (Bundes<strong>an</strong>stalt für<br />

Geowissenschaften und Rohst<strong>of</strong>fe), 1991. Symbolschlüssel Geologie. E.<br />

Schweizerbart`scheVerlagsbuchh<strong>an</strong>dlung, Stuttgart.<br />

Olphen <strong>an</strong>d Fripiat, 1979. Data H<strong>an</strong>dbooks for Clay Minerals <strong>an</strong>d Non-metallic Minerals.<br />

Ot<strong>an</strong>do, S. W., 1996. Regional <strong>geological</strong> evaluation <strong>of</strong> Nairobi – Kiambu areas<br />

(MSc thesis - unpublished).<br />

Petley, D. J., 1966. The shear strength <strong>of</strong> soils at large strains. PhD Thesis, University <strong>of</strong> London.<br />

Pike, D. C., 1973. Shearbox tests on graded aggregates. TRRL Report LR 584, Tr<strong>an</strong>sport <strong>an</strong>d Road<br />

Research Laboratoy, Crowthorne, Berks.<br />

-----, Acott, S. M.<strong>an</strong>d Leech, R. M., (1977). Sub-base stability: A shearbox test compared with<br />

other prediction methods. Tr<strong>an</strong>sport <strong>an</strong>d Road Research Laboratory, Report No. LR 785.<br />

Prior, G. T, 1903. Contributions to the petrology <strong>of</strong> British East Africa. Min. Mag. 13, 228-263.


216<br />

Pulfrey, W., 1949. Ijolitic rocks near Homa Bay, Western Kenya. Q. J. geol. Soc. London. 105,<br />

425-454.<br />

Richwien, W. und Lesny, K., 2000. Bodenmech<strong>an</strong>isches Praktikum: Auswahl und<br />

Anwendung von bodenmech<strong>an</strong>ischen Laborversuchen. Verlag Glückauf GmbH, Essen.<br />

Rösler, H. J., 1980. Lehrbuch der Mineralogie. VEB Deutscher Verlag für Grundst<strong>of</strong>findustrie, Leipzig.<br />

Rowe, P. W., 1966. A new consolidation cell. Geotechnique, Vol. 16, No. 2.<br />

-----, 1972. The relev<strong>an</strong>ce <strong>of</strong> soil fabric to site investigation practice. 12 th R<strong>an</strong>kine lecture,<br />

Geotechnique, Vol. 22, No. 2.<br />

Saggerson, E. P., Unpublished. Chemical <strong>an</strong>alyses <strong>of</strong> igneous, metamorphic <strong>an</strong>d sedimentary<br />

rocks <strong>an</strong>d natural gas from Kenya. Bull. geol. Surv. Kenya.<br />

-----, Joubert, P., McCall, G. J. G. <strong>an</strong>d Williams, L. A. J., 1960. Cross-folding <strong>an</strong>d refolding<br />

in the Basement System <strong>of</strong> Kenya Colony. 21 st Session <strong>of</strong> the International Geological Congress,<br />

Sect. 18, 335-346.<br />

------ <strong>an</strong>d Baker, B. H., 1965. Post-Jurassic erosion surfaces <strong>an</strong>d their relationship to Rift<br />

structure. Q. J. geol. Soc. London 121, 51-72.<br />

-----, 1991. Geology <strong>of</strong> the Nairobi area. Rep. Mines <strong>an</strong>d geol. Dept., Kenya 98.<br />

Sahama, Th. G., 1952. Leucite, potash nepheline <strong>an</strong>d clinopyroxene from south-west Ug<strong>an</strong>da<br />

<strong>an</strong>d Belgi<strong>an</strong> Congo. Am. J. Sci. Bowen Volume.<br />

Sattler, C. D., 2000. Röntgendiffraktometrie in der Untersuchung von Sedimenten. S 4822<br />

Sedimentpetrographische Arbeitsmethoden III, TU Clausthal.<br />

Schmertm<strong>an</strong>n, J. H., 1953. Estimating the true consolidation behaviour <strong>of</strong> clay from<br />

laboratory test results. Proc. ASCE, Vol. 79, Separate No. 3111.<br />

-----, 1954. The undisturbed consolidation behaviour <strong>of</strong> clay. Tr<strong>an</strong>s. ASCE, Vol. 120, Paper 2775.<br />

Scott, C. R., 1974. An introduction to Soil Mech<strong>an</strong>ics. Applied Science Publishers.<br />

Scott, R. M.,1963. The soils <strong>of</strong> the Nairobi – Thika – Yatta – Machakos area. Rep. Dept.<br />

Agriculture, Kenya (not numbered).<br />

Seismology Section, Department <strong>of</strong> Geology, University <strong>of</strong> Nairobi, 2001. Reports on seismic<br />

actitvities in Nairobi <strong>an</strong>d the Rift region. Unpublished.<br />

Sh<strong>an</strong>d, S. J., 1937. Rocks <strong>of</strong> the Kenya Rift Valley. Geol. Mag. 74, 262-271.<br />

Sherwood, P. T., 1967. Classification tests on Afric<strong>an</strong> red <strong>clays</strong> <strong>an</strong>d Keuper Marl. Q. J. Eng.<br />

Geol. 1, 47-53.<br />

Sikes, H.L., 1939. Notes on the geology <strong>of</strong> the country surrounding Nairobi. Report (not<br />

numbered), Geol. Surv. Kenya.<br />

------,1934. The undeground water resources <strong>of</strong> Kenya Colony. Crown Agents, London.<br />

------, H.L., 1926. The structure <strong>of</strong> the eastern fl<strong>an</strong>k <strong>of</strong> the Great Rift Valley near Nairobi.<br />

Geogr. J., 68: 385-402.<br />

Skempton, A. W., 1944. Notes on the compressibility <strong>of</strong> <strong>clays</strong>. Q. J. Geol. Soc., London,<br />

Vol. C.


217<br />

------,1964. Long term stability <strong>of</strong> clay slopes. Fourth R<strong>an</strong>kine Lecture,<br />

Geotechnique, Vol. 14, No. 2.<br />

----- <strong>an</strong>d Northey, R. D., 1953. The sensitivity <strong>of</strong> <strong>clays</strong>. Geotechnique, Vol. 3, No. 1.<br />

-----, <strong>an</strong>d Bjerum, L., 1957. A contribution to the settlement <strong>an</strong>alysis <strong>of</strong> foundations on clay.<br />

Geotechnique, Vol. 7, p. 168.<br />

----- <strong>an</strong>d La Rochelle, P., 1965. The Bradwell slip ; A short term failure in<br />

London clay. Geotechnique, Vol. 15, No. 3.<br />

Smith, W. Campbell, 1931. A classification <strong>of</strong> some rhyolites, trachytes <strong>an</strong>d phonolites from<br />

part <strong>of</strong> Kenya Colony, with a note on some associated basaltic rocks. Q. J. geol. Soc.<br />

London 94, 249-254 <strong>an</strong>d 507-553.<br />

-----, 1950. Note on the Kapiti Phonolites <strong>an</strong>d kenytes <strong>of</strong> Kenya Colony. Bull. Br. Mus.<br />

(Nat. Hist.) Min. 1, 1-15.<br />

Smith, J. V. <strong>an</strong>d Sahama Th. G., 1954. Determination <strong>of</strong> the composition <strong>of</strong> Nephelines by <strong>an</strong><br />

X-ray method. Min. Mag. 30, 439-449.<br />

Sowers, G.F., 1994. Geotechnical issues in site selection <strong>an</strong>d project development. Bull. Assoc.<br />

Eng. Geol., 31: 209-216.<br />

Spencer, E., 1937. The potash-soda feldspars. Min. Mag. 24, 453.<br />

Stephen, I., Bellis, F. <strong>an</strong>d Muir, A., 1956. Gilgai phenomena in <strong>tropical</strong> black <strong>clays</strong> <strong>of</strong><br />

Kenya. J. Soil Sci. 7, 1-9.<br />

Stokes, Sir George G., 1891. Mathematical <strong>an</strong>d Physical Paper III. Cambridge Univ. Press.<br />

Suess, E, 1892. Die Bruche des östlichen Afrika. Denkschr. Akad. Wiss., Wien Math-nat. 58, 555-584.<br />

Survey <strong>of</strong> Kenya, 1991. Atlas <strong>of</strong> Kenya, 3 rd ed.<br />

Symons, I. F., (1968). The application <strong>of</strong> residual shear strength to the design <strong>of</strong> cuttings in<br />

overconsolidated fissured <strong>clays</strong>. Report No. LR 227, Tr<strong>an</strong>sport <strong>an</strong>d Road Research Laboratoy,<br />

Crowthorne, Berks.<br />

Taylor, D. W., 1948. Fundamentals <strong>of</strong> Soil Mech<strong>an</strong>ics. Wiley, New York .<br />

-------, 1942. Research on consolidation <strong>of</strong> <strong>clays</strong>. M.I.T., Dept. <strong>of</strong> Civ. <strong>an</strong>d S<strong>an</strong>it.<br />

Eng., No. 82.<br />

Terzaghi, K., 1925. Erdbaumech<strong>an</strong>ik auf bodenphysikalischer Grundlage. Deuticke, Wien.<br />

-----, 1934. Die Ursachen der Schiefstellung des Turmes von Pisa. Der Bauingenieur, Berlin.<br />

-----, 1939. Soil Mech<strong>an</strong>ics- a new chapter in <strong>engineering</strong> science. James Forrest Lecture,<br />

J. Inst. Civ. Eng., London, Vol. 12, No. 7.<br />

-----, 1943. Theoretical Soil Mech<strong>an</strong>ics. Wiley, New York.<br />

-----, 1958. Design <strong>an</strong>d perform<strong>an</strong>ce <strong>of</strong> the Sasumua Dam. Proc. Inst. geol. Engrs. 9, 369-<br />

396.<br />

------ <strong>an</strong>d Peck, R. B., 1948. Soil Mech<strong>an</strong>ics in Engineering Practice. Wiley, New York.<br />

------ <strong>an</strong>d Peck, R. B., 1967. Soil Mech<strong>an</strong>ics in Engineering Practice. 2 nd edition, Wiley, New<br />

York.


218<br />

Thomson Scientific Instruments Pty Ltd. & Fa.Getac, 1995. Hardware for High Perform<strong>an</strong>ce<br />

X-ray Micro-<strong>an</strong>alysis (EDX:WINEDS) <strong>an</strong>d S<strong>of</strong>tware (Wineds 3.0; Edisson 32) for<br />

Element Analysis.<br />

Tilford, N.R., 1994. Site selection: past <strong>an</strong>d present. Bull. Assoc. Eng. Geol., 31:157-169.<br />

Tilley, C.E., 1956. Nepheline-alkali feldspar paragenesis. Am. J. Sci. 252.<br />

Tschebotari<strong>of</strong>f, G. P., (1951). Soil Mech<strong>an</strong>ics, Foundations <strong>an</strong>d Earth Structures.<br />

McGraw-Hill, New York.<br />

Tsiambaos, G. <strong>an</strong>d Tsaligopoulos, CH., 1995. A proposed method <strong>of</strong> estimating the swelling<br />

characteristics <strong>of</strong> soils (examples from Greece). Bull. IAEG, 52: 109-115, Paris.<br />

Verstappen, H.Th., 1980. Applied Geomorphology. Elsevier, Amsterdam, 437 pp.<br />

Walker, E. E, 1903. Report on the geology <strong>of</strong> the East Africa Protectorate. Africa 11 (Cd. 1769).<br />

Wetherill, G. B., 1966. Sequential methods in statistics. London: Methuen.<br />

Williams, L. A. J, 1967. In Nairobi: city <strong>an</strong>d region. 1-13 (W. T. W. Morg<strong>an</strong> ed). Oxford Univ. Press,<br />

London.<br />

Willis, B, 1936. East Afric<strong>an</strong> plateaus <strong>an</strong>d rift valleys. Carnegie Institute, Washington.<br />

Wood, D. M. <strong>an</strong>d Wroth, C. P., 1978. The use <strong>of</strong> the cone penetrometer to determine the<br />

plastic limit <strong>of</strong> soils. Ground Engineering, Vol. 11, No. 3.


Appendix A: Oedometer consolidation tests<br />

Complete results for the loading/ unloading stages <strong>of</strong> black <strong>clays</strong> <strong>an</strong>d red soils<br />

Table A1. Consolidation parameters for black <strong>clays</strong>, Sample No. SA2-70cm.<br />

Calculation sheet Sample No.: SA2-70cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 0,709 - - - - - -<br />

100 0,043 0,307 0,704 0,000052 0,031 1,00 5,080 4,837E-11 0,30 0,070 0,0002<br />

200 0,408 2,914 0,659 0,000446 0,262 9,50 0,519 4,210E-11 0,004<br />

400 0,935 6,679 0,595 0,000322 0,194 28,00 0,165 9,914E-12 0,007<br />

800 1,556 11,114 0,519 0,000190 0,119 60,00 0,070 2,597E-12 0,009<br />

1600 2,292 16,371 0,429 0,000112 0,074 60,00 0,063 1,448E-12 0,011<br />

2400 2,708 19,343 0,379 0,000063 0,044 95,00 0,036 4,983E-13 0,007<br />

1600 2,601 18,579 0,392<br />

400 1,942 13,871 0,472<br />

100 1,235 8,821 0,558<br />

Table A2. Consolidation parameters for black <strong>clays</strong>, Sample No. SA37-50cm.<br />

Calculation sheet<br />

Sample No.: SA37-50cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 1,098 - - - - - -<br />

25 0,258 1,843 1,060 0,001547 0,737 2,90 1,725 3,942E-10 0,003<br />

50 0,813 5,807 0,977 0,003328 1,615 7,00 0,673 3,372E-10 0,40 0,08 0,004<br />

100 1,511 10,793 0,872 0,002092 1,059 12,00 0,357 1,172E-10 0,008<br />

200 2,283 16,307 0,756 0,001157 0,618 27,00 0,141 2,703E-11 0,010<br />

400 3,085 22,036 0,636 0,000601 0,342 45,00 0,074 7,849E-12 0,015<br />

800 3,875 27,679 0,518 0,000296 0,181 55,00 0,052 2,935E-12 0,017<br />

1600 4,290 30,643 0,455 0,000078 0,051 95,00 0,027 4,275E-13<br />

100 3,700 26,429 0,544<br />

25 3,400 24,286 0,589<br />

12 3,087 22,050 0,636<br />

Table A3. Consolidation parameters for black <strong>clays</strong>, Sample No. SB1-70cm.<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 1,045 - - - - - -<br />

50 0,259 1,850 1,007 0,000757 0,370 0,48 10,421 1,195E-09 0,0002<br />

100 0,770 5,500 0,933 0,001493 0,744 2,00 2,364 5,451E-10 0,002<br />

200 1,585 11,321 0,814 0,001191 0,616 1,80 2,375 4,535E-10 0,3 0,043 0,004<br />

400 2,320 16,571 0,706 0,000537 0,296 11,00 0,343 3,148E-11 0,005<br />

800 2,725 19,464 0,647 0,000148 0,087 18,00 0,190 5,113E-12 0,006<br />

1600 3,303 23,593 0,563 0,000106 0,064 40,00 0,078 1,558E-12 0,012<br />

2400 3,638 25,986 0,514 0,000061 0,039 92,00 0,031 3,802E-13 0,008<br />

1600 3,605 25,750 0,518<br />

400 3,321 23,721 0,560<br />

50 2,666 19,043 0,656


Table A4. Consolidation parameters for black <strong>clays</strong>, Sample No. SB42-50cm.<br />

Calculation sheet Sample No.: SB42-50cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 0,732 - - - - - -<br />

100 1,123 8,021 0,593 0,001389 0,80 9,50 0,494 1,229E-10 0,005<br />

200 2,041 14,579 0,480 0,001136 0,71 20,00 0,200 4,430E-11 0,009<br />

400 3,098 22,129 0,349 0,000654 0,44 24,00 0,142 1,939E-11 0,024<br />

800 4,075 29,107 0,228 0,000302 0,22 50,00 0,056 3,916E-12 0,34 0,03 0,017<br />

1600 4,972 35,514 0,117 0,000139 0,11 70,00 0,033 1,168E-12 0,021<br />

2400 5,404 38,600 0,064 0,000067 0,06 95,00 0,021 3,941E-13 0,009<br />

1600 5,338 38,129 0,072<br />

400 4,732 33,800 0,147<br />

100 4,155 29,679 0,218<br />

25 3,677 26,264 0,277<br />

12 3,075 21,964 0,352<br />

Table A5. Consolidation parameters for black <strong>clays</strong>, Sample No. SC17-50cm<br />

Calculation sheet Sample No.: SC17-50cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 0,728 - - - - - - - -<br />

100 0,244 1,743 0,698 0,000301 0,17 9,00 0,556 3,006E-11 0,006<br />

200 1,067 7,621 0,596 0,001016 0,60 30,00 0,154 2,862E-11 0,007<br />

400 1,999 14,279 0,481 0,000575 0,36 18,00 0,225 2,508E-11 0,009<br />

800 2,889 20,636 0,371 0,000275 0,19 50,00 0,069 3,991E-12 0,016<br />

1600 3,962 28,300 0,239 0,000166 0,12 65,00 0,045 1,674E-12 0,44 0,040 0,020<br />

2400 4,386 31,329 0,187 0,000065 0,05 120,00 0,021 3,424E-13 0,012<br />

1600 4,317 30,836 0,195<br />

400 3,810 27,214 0,258<br />

70 3,043 21,736 0,352<br />

Table A6. Consolidation parameters for black <strong>clays</strong>, Sample No. SC29-50cm.<br />

Consolidation test calculation sheet Sample No: SC29-50cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 0,901 - - - - - -<br />

100 0,466 3,329 0,838 0,000633 0,33 2,50 1,971 2,034E-10 0,003<br />

200 1,061 7,579 0,757 0,000808 0,44 5,50 0,828 1,129E-10 0,004<br />

400 1,791 12,793 0,658 0,000496 0,28 15,00 0,274 2,397E-11 0,009<br />

800 2,553 18,236 0,555 0,000259 0,16 35,00 0,104 5,027E-12 0,013<br />

1600 3,418 24,414 0,437 0,000147 0,09 38,00 0,083 2,431E-12 0,014<br />

2400 3,807 27,193 0,384 0,000066 0,05 80,00 0,035 4,995E-13 0,38 0,065 0,008<br />

1600 3,735 26,679 0,394<br />

400 3,256 23,257 0,459<br />

50 2,585 18,464 0,550


Table A7. Consolidation parameters for red soils, Sample No. RD1-30cm<br />

Consolidation test<br />

calculation sheet Sample No.: RD1-30cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 1,560 - - - - - - - -<br />

12 1,205 8,607 1,340 0,018365 7,17 0,30 15,56 3,459E-08 0,002<br />

25 1,537 10,979 1,279 0,004671 2,00 0,35 11,85 7,331E-09 0,003<br />

50 1,770 12,643 1,237 0,001705 0,75 1,05 3,77 8,750E-10 0,003<br />

100 2,338 16,700 1,133 0,002078 0,93 1,25 2,97 8,547E-10 0,004<br />

200 2,990 21,357 1,014 0,001192 0,56 1,50 2,23 3,860E-10 0,004<br />

400 3,720 26,571 0,880 0,000668 0,33 1,60 1,84 1,892E-10 0,004<br />

800 4,490 32,071 0,739 0,000352 0,19 1,85 1,38 7,988E-11 0,004<br />

1600 5,174 36,957 0,614 0,000156 0,09 2,20 0,99 2,769E-11 0,003<br />

2400 5,523 39,450 0,550 0,000080 0,05 3,90 0,50 7,646E-12 0,42 0,030 0,003<br />

1600 5,503 39,307 0,554<br />

400 5,410 38,643 0,571<br />

100 5,323 38,021 0,587<br />

25 5,185 37,036 0,612<br />

12 5,092 36,371 0,629<br />

Table A8. Consolidation parameters for red soils, Sample No. RD1-100cm.<br />

Consolidation test<br />

calculation sheet Sample No.: RD1-100cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 1,607 - - - - - - - -<br />

12 0,513 3,664 1,511 0,007961 3,05 0,16 30,69 2,905E-08 0,003<br />

25 1,029 7,350 1,415 0,007391 2,94 0,98 4,67 4,258E-09 0,005<br />

50 1,816 12,971 1,269 0,005862 2,43 1,00 4,11 3,094E-09 0,007<br />

100 2,781 19,864 1,089 0,003594 1,58 1,11 3,21 1,575E-09 0,008<br />

200 3,787 27,050 0,902 0,001873 0,90 1,40 2,13 5,928E-10 0,008<br />

400 4,693 33,521 0,733 0,000844 0,44 1,50 1,65 2,270E-10 0,006<br />

800 5,462 39,014 0,590 0,000358 0,21 1,75 1,18 7,574E-11 0,005<br />

1600 6,095 43,536 0,472 0,000147 0,09 2,10 0,84 2,404E-11 0,4 0,030 0,004<br />

2400 6,396 45,686 0,416 0,000070 0,05 3,80 0,41 6,071E-12 0,004<br />

1600 6,383 45,593 0,418<br />

400 6,288 44,914 0,436<br />

100 6,164 44,029 0,459<br />

25 6,033 43,093 0,484<br />

12 5,952 42,514 0,499


Table A9. Consolidation parameters for red soils, Sample No. RD1-400cm.<br />

Consolidation test<br />

calculation sheet Sample No.: RD1-400cm<br />

P (kPa) DH (mm) E (%) e av ( m²/kN) mv (m²/MN) t50 (min) Cv (m²/year) K (m/s) Cc Cs Cα<br />

0 0,000 0,000 1,163 - - - - - - -<br />

12 0,233 1,664 1,127 0,003000 1,39 0,33 15,42 6,630E-09 0,003<br />

25 0,432 3,086 1,096 0,002365 1,11 0,40 12,14 4,185E-09 0,002<br />

50 0,634 4,529 1,065 0,001248 0,60 0,56 8,42 1,554E-09 0,002<br />

100 0,940 6,714 1,018 0,000945 0,46 0,85 5,34 7,580E-10 0,003<br />

200 1,374 9,814 0,951 0,000670 0,33 0,90 4,77 4,909E-10 0,005<br />

400 1,986 14,186 0,856 0,000473 0,24 1,10 3,59 2,695E-10 0,006<br />

800 2,744 19,600 0,739 0,000293 0,16 1,30 2,71 1,324E-10 0,007<br />

1600 3,493 24,950 0,623 0,000145 0,08 1,80 1,71 4,410E-11 0,39 0,020 0,006<br />

2400 3,878 27,700 0,564 0,000074 0,05 2,50 1,11 1,571E-11 0,006<br />

1600 3,863 27,593 0,566<br />

400 3,779 26,993 0,579<br />

100 3,675 26,250 0,595<br />

25 3,528 25,200 0,618<br />

12 3,249 23,207 0,661


Results <strong>of</strong> swelling tests on black <strong>clays</strong><br />

Appendix B: Swelling tests<br />

Swelling S (mm) is expressed as percentage <strong>of</strong> ultimate swelling Smax (mm) at zero loading.<br />

Table B1<br />

Table B2<br />

Swelling test: Sample No. SA2/70cm Swelling test: Sample No.SB1/70cm<br />

Initial sample size<br />

Initial sample size<br />

Diameter (mm) 71,4 Diameter (mm) 71,4<br />

Height Ho (mm) 11 Height Ho (mm) 11<br />

X-Sectional area (m²) 0,004 X-Sectional area (m²) 0,004<br />

Swelling pressure load (Kg) 33 Swelling pressure load (Kg) 15<br />

Swelling pressure SP (kPa) 82,46 Swelling pressure SP (kPa) 37,48<br />

Maximum swelling Smax (mm) 1,036 Maximum swelling Smax (mm) 0,425<br />

Load Swelling<br />

(kg) (kPa) (mm) S (% ) Load Swelling<br />

33 82,46 0 0 (kg) P (kPa) (mm) S (%)<br />

16,5 41,23 0,106 10 15 37,48 0 0<br />

8 19,99 0,253 24 7,5 18,74 0,054 13<br />

4 10,00 0,321 31 4 10,00 0,126 30<br />

2 5,00 0,539 52 2 5,00 0,197 46<br />

1 2,50 0,725 70 1 2,50 0,279 66<br />

0 0,00 1,036 100 0 0,00 0,425 100<br />

Table B3<br />

Table B4<br />

Swelling test: Sample No. SC25/50cm Swelling test: Sample No. SC29/50cm<br />

Initial sample size<br />

Initial sample size<br />

Diameter (mm) 71,4 Diameter (mm) 71,4<br />

Height Ho (mm) 11 Height Ho (mm) 11<br />

X-Sectional area (m²) 0,004 X-Sectional area (m²) 0,004<br />

Swelling pressure load (Kg) 14,5 Swelling pressure load (Kg) 18<br />

Swelling pressure SP (kPa) 36,23 Swelling pressure SP (kPa) 44,98<br />

Maximum swelling Smax (mm) 0,475 Maximum swelling Smax (mm) 0,493<br />

Load Swelling Load Swelling<br />

(kg) P (kPa) (mm) S (%) (kg) P (kPa) (mm) S (%)<br />

14,5 36,23 0,000 0 18 44,98 0,000 0<br />

7 17,49 0,066 14 9 22,49 0,061 12<br />

3,5 8,75 0,160 34 4,5 11,24 0,149 30<br />

2 5,00 0,260 55 2 5,00 0,266 54<br />

1 2,50 0,347 73 1 2,50 0,356 72<br />

0 0,00 0,475 100 0 0,00 0,493 100


Results <strong>of</strong> swelling tests on black <strong>clays</strong><br />

Appendix C: Swelling tests<br />

Swelling S (mm) is expressed as percentage <strong>of</strong> initial specimen thickness Ho (mm).<br />

Table C1<br />

Table C2<br />

Swelling test: Sample No. SA2/70cm Swelling test: Sample No. SB1/70cm<br />

Initial sample size<br />

Initial sample size<br />

Diameter (mm) 71,4 Diameter (mm) 71,4<br />

Height Ho (mm) 11 Height Ho (mm) 11<br />

X-Sectional area (m²) 0,004 X-Sectional area (m²) 0,004<br />

Swelling pressure load (Kg) 33 Swelling pressure load (Kg) 15<br />

Swelling pressure SP (kPa) 82,46 Swelling pressure SP (kPa) 37,48<br />

Maximum swelling Smax (mm) 1,036 Maximum swelling Smax (mm) 0,425<br />

Load<br />

Swelling<br />

(kg) (kPa) (mm) % Load Swelling<br />

33 82,46 0,000 0,00 (kg) (kPa) (mm) %<br />

16,5 41,23 0,106 0,96 15 37,48 0,000 0,00<br />

8 19,99 0,253 2,30 7,5 18,74 0,054 0,49<br />

4 10,00 0,321 2,92 4 10,00 0,126 1,15<br />

2 5,00 0,539 4,90 2 5,00 0,197 1,79<br />

1 2,50 0,725 6,59 1 2,50 0,279 2,54<br />

0 0,00 1,036 9,42 0 0,00 0,425 3,86<br />

Table C3<br />

Table C4<br />

Swelling test: Sample No. SC25/50cm Swelling test: Sample No. SC29/50cm<br />

Initial sample size<br />

Initial sample size<br />

Diameter (mm) 71,4 Diameter (mm) 71,4<br />

Height Ho (mm) 11 Height Ho (mm) 11<br />

X-Sectional area (m²) 0,004 X-Sectional area (m²) 0,004<br />

Swelling pressure load (Kg) 14,5 Swelling pressure load (Kg) 18<br />

Swelling pressure SP (kPa) 36,23 Swelling pressure SP (kPa) 44,98<br />

Maximum swelling Smax (mm) 0,475 Maximum swelling Smax (mm) 0,493<br />

Load Swelling Load Swelling<br />

(kg) (kPa) (mm) % (kg) (kPa) (mm) %<br />

14,5 36,23 0 0,00 18 44,98 0 0,00<br />

7 17,49 0,066 0,60 9 22,49 0,061 0,55<br />

3,5 8,75 0,16 1,45 4,5 11,24 0,149 1,35<br />

2 5,00 0,26 2,36 2 5,00 0,266 2,42<br />

1 2,50 0,347 3,15 1 2,50 0,356 3,24<br />

0 0,00 0,475 4,32 0 0,00 0,493 4,48


Appendix D<br />

Distribution/ variation <strong>of</strong> index <strong>an</strong>d <strong>engineering</strong> properties <strong>of</strong><br />

black <strong>clays</strong> in Nairobi area


16000<br />

14000<br />

12000<br />

10000<br />

West<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000 22000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D1 Distribution/ depth variation (m) <strong>of</strong> soils in Nairobi area (Ot<strong>an</strong>do S. W., 2003)<br />

Dist<strong>an</strong>ce (m)


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D2(a) Distribution/ moisture (Wn%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

West<br />

1000<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D2(b) Distribution/ moisture (Wn%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

1000<br />

Dist<strong>an</strong>ce (m)<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D3(a) Distribution/ liquid limit (LL%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less<br />

th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

1000<br />

Dist<strong>an</strong>ce (m)<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D3(b) Distribution/ liquid limit (LL%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater<br />

th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D4(a) Distribution/ plasticity index (PI%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D4(b) Distribution/ plasticity index (PI%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D5(a) Distribution/ linear shrinkage (LS%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D5(b) Distribution/ linear shrinkage (LS%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D6(a) Distribution/ free swell (FS%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D6(b) Distribution/ free swell (FS%) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D7(a) Distribution/ fines fraction (%fines) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D7(b) Distribution/ fines fraction (%fines) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D8(a) Distribution/ clay fraction (%clay) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D8(b) Distribution/ clay fraction (%clay) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D9(a) Distribution/ <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (phi °) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m<br />

(Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D9(b) Distribution/ <strong>an</strong>gle <strong>of</strong> shear resist<strong>an</strong>ce (phi °) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m<br />

(Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D10(a) Distribution/ apparent cohesion (c' kPa)) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D10(b) Distribution/ apparent cohesion (c' kPa)) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D11(a) Distribution/ activity level (A) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths less th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


4000<br />

3000<br />

2000<br />

Dist<strong>an</strong>ce (m)<br />

1000<br />

West<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

South<br />

Dist<strong>an</strong>ce (m)<br />

Fig. D11(b) Distribution/ activity level (A) variation <strong>of</strong> black <strong>clays</strong> in Nairobi area for depths greater th<strong>an</strong> 0,5m (Ot<strong>an</strong>do S. W., 2003).


Appendix E<br />

Geotechnical soil map <strong>of</strong> Nairobi area, Kenya


15000<br />

10000<br />

5000<br />

Dist<strong>an</strong>ce (m)<br />

0<br />

0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000 22000<br />

m<br />

Fig. E1 Geotechnical soils map <strong>of</strong> Nairobi area, Kenya (prepared by Ot<strong>an</strong>do, S. W., 2003).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!