You are on page 1of 201

IMPERIAL COLLEGE LONDON

Faculty of Engineering

Department of Civil and Environmental Engineering

Structural Controls on the


Hydrogeology of Malta

Christian Schembri
September 2014

Submitted in fulfilment of the requirements for the MSc and the Diploma of Imperial College London

DECLARATION OF OWN WORK

Declaration:

This submission is my own work. Any quotation from, or description of, the
work of others is acknowledged herein by reference to the sources, whether
published or unpublished.

Signature : ___________________________________

Page | ii

To my beloved wife Rosanne

Page | iii

The research work disclosed in this publication is partially funded by the Master it!
Scholarship Scheme (Malta). This Scholarship is part-financed by the European
Union European Social Fund (ESF) under Operational Programme II

Cohesion

Policy 2007-2013, Empowering People for More Jobs and a Better Quality Of Life.

Operational Programme II Cohesion Policy 2007-2013


Empowering People for More Jobs and a Better Quality of
Life
Scholarship part-financed by the European Union
European Social Fund (ESF)
Co-financing rate: 85% EU Funds;15% National Funds
Investing in your future

Page | iv

Acknowledgements
First and foremost I would like to thank my wife Rosanne for her emotional support
over the last year. The past year has been a challenge for both of us and thus I
would also like to thank our extended family members and friends for their support.
Special thanks go to Dr Clark Fenton of the Imperial College London for his help,
inspiration and guidance over the last year and to the staff at the Geotechnics
Department.
I would also like to thank Dr Martyn Pedley and Dr Adrian Butler for their replies for
my queries. I would like to express my gratitude to Adrian Mifsud for his assistance
during my field trip and interest. Thanks go also to Solidbase Laboratories Ltd and
Joe Bugeja for allowing me access to certain documents.
Thanks go to Roderick Vella from the Transport and Infrastructure Ministry of Malta
for providing me with useful contacts during data collection and Manuel Sapiano for
providing me with access to past reports and for his prompt replies to my queries.
My gratitude goes also to my fellow students who have been ideal colleagues and
friends throughout the year.

Page | v

Executive Summary
The main aim of this study is to identify and understand structural controls on the
hydrogeology of Malta.
An introduction chapter provides a general geologic context of Malta, followed by a
detailed aim and an overall view of the study. This is followed by an extensive
literature review providing a sound geologic and hydrogeologic background.
The Maltese Islands are located in the foreland of the Apennine-Maghrebian thrust
and fold belt and are affected by an extensional tectonic environment. An onshore
expressed of this are the horst and graben structures widely observable in the range
between the South of Gozo fault (or Il-Qala fault) and the Victoria fault. Tectonics
controlled the sedimentation processes. Similar processes but more pronounced
were taking place at the offshore regional grabens of the Pantelleria rift and the
North Gozo graben. Higher extensional strains are reported in the latter basins than
what is reported for onshore Malta. The uplift of the Maltese Islands occurred
during some period stretching between the late Messinian to the mid-Pliocene. The
geologic formations of Malta consist in sedimentary marine carbonates deposited at
shallow sea depths with the highest sea depths estimated not to be greater than
250 metres. These include from top, the Upper Coralline Limestone (UCL),
Greensand, Blue Clay (BC), Globigerina Limestone (GL) and Lower Coralline
Limestone (LCL).
Malta has two main types of aquifers being the perched aquifer on top of the BC
and the mean sea level aquifer which is predominantly hosted by the LCL. The
aquifers are dual-porosity as flow may take place through the rock matrix and the
fracture network. Faults may have two contrasting effects. They may provide a seal
or increase the density of the fracture network.
The main data collection process included a one-week field trip during which a
geomorphologic site reconnaissance exercise was carried out and discontinuity scan
line data was collected. Previous data was also acquired and is re-interpreted and
Page | vi

used. This data set includes investigation borehole logs, well pumping tests
determining aquifer transmissivity and potentiometric data. Previous data is
generally limited to a regional context, incomplete and does not satisfy directly the
scope of this study. Predictive data from previous field studies is used to augment
our understanding of the effects of faults on the hydraulic conductivity of rock
masses. A detailed observations and analysis chapter is presented. The main
findings are summarised as follows:

The hypothesis that joints are closely linked to the latest rift tectonics of
Malta is presented. Evidence includes similar strikes and dip angles of the
most occurring joints that closely resemble the ENE-WSW and NW-SE
trending faults and the wider apertures of these joints.

Exceptions to the above rule may be present as is observed at the site of


Birkirkara where the main trending joint set J8 shows strike similarity to the
ENE-WSW trending faults but occurs at a much shallower dip. This probably
is related to another structure which in literature is described as the uparching of the LGL prior to the rifting process.

From observations it is noted that karst development differs between


formations or facies that exhibit variability in hydraulic permeability due to
grain size distribution and fracturing. In addition observations highlight that
fluid conducting boundaries can result between layers of different grain size
distributions.

A plot of spatial transmissivity in relation to distance from faults, although


from low resolution and an incomplete data set, provides encouraging
indications for future research as a certain degree of correlation between
the two seems plausible.

Page | vii

Table of Contents
Acknowledgements ...................................................................................................... v
Executive Summary ..................................................................................................... vi
Table of Contents .......................................................................................................viii
Tables ...........................................................................................................................xi
Figures .........................................................................................................................xii
1

Introduction .......................................................................................................... 1
1.1

Scope of Work ............................................................................................... 3

1.2

Overview of Work.......................................................................................... 4

Geologic and Hydrogeologic Background of Malta .............................................. 6


2.1

2.1.1.

Tectonic Setting...................................................................................... 6

2.1.2.

Tectonic History Debates ..................................................................... 11

2.1.3.

Onshore Structural Geology of Malta .................................................. 12

2.1.4.

Main Stratigraphical Units ................................................................... 14

2.2

Geologic Background ..................................................................................... 6

Hydrogeologic Background ......................................................................... 17

2.2.1.

Basic Hydroclimatological data ............................................................ 17

2.2.2.

Water Balance ...................................................................................... 18

2.2.3.

Hydrogeological Setting ....................................................................... 18

2.2.4.

Effects of faults on fluid flow ............................................................... 20

2.2.5.

Aquifer hydraulic properties ................................................................ 22

2.2.6.

Geochemical Studies ............................................................................ 23

Methods .............................................................................................................. 25
3.1

Geomorphologic Site Reconnaissance ........................................................ 26

3.2

Dip (angles) and dip directions of discontinuities ....................................... 27

3.3

Other discontinuities characteristics........................................................... 30

3.3.1.

Relative hydraulic conductivity ............................................................ 30

3.4

Transmissivity .............................................................................................. 31

3.5

Potentiometry ............................................................................................. 32

3.6

Control of fault parameters on hydraulic properties .................................. 33

3.6.1.

Displacements along fault lengths ....................................................... 33

3.6.2.

Length and termination points of faults .............................................. 35


Page | viii

3.6.3.

Fault architecture at a cross-section .................................................... 36

3.6.4.

Strata thickness and properties ........................................................... 37

Observations and Analysis .................................................................................. 38


4.1

Geomorphologic Site Reconnaissance ........................................................ 38

4.1.1.

Qammiegh ............................................................................................ 38

4.1.2.

L-Imgiebah Bay ..................................................................................... 41

4.1.3.

Fomm ir-Rih Bay ................................................................................... 45

4.1.4.

Wied il-Ghasel ...................................................................................... 48

4.1.5.

Gharghur .............................................................................................. 48

4.1.6.

St. Georges Bay ................................................................................... 51

4.1.7.

Msida .................................................................................................... 52

4.1.8.

Xghajra ................................................................................................. 54

4.1.9.

Munxar ................................................................................................. 56

4.2

Inferring Contacts from Boreholes .............................................................. 58

4.2.1.
4.3

Dip (angles) and dip directions of discontinuities ....................................... 59

4.3.1.

Fomm ir-Rih Bay ................................................................................... 59

4.3.2.

St. Georges Bay ................................................................................... 60

4.3.3.

Msida .................................................................................................... 61

4.3.4.

Xghajra ................................................................................................. 61

4.3.5.

Birkirkara .............................................................................................. 62

4.4

UCL/BC ................................................................................................. 58

Other discontinuities characteristics........................................................... 63

4.4.1.

Aperture ............................................................................................... 63

4.4.2.

Persistence ........................................................................................... 64

4.4.3.

Relative hydraulic conductivity (K)....................................................... 66

4.5

Transmissivity .............................................................................................. 66

4.6

Potentiometry ............................................................................................. 71

Discussion ........................................................................................................... 72
5.1

Geomorphologic Site Reconnaissance ........................................................ 72

5.1.1.

Stratal Dip of BC ................................................................................... 72

5.1.2.

Flow indications from karst erosion..................................................... 73

5.1.3.

Sedimentation processes ..................................................................... 74

5.1.4.

Calcite Deposition ................................................................................ 74

Page | ix

5.1.5.
5.2

Style of faulting .................................................................................... 75

Discontinuity data ....................................................................................... 75

5.2.1.

Jointing link with the rifting tectonics of Malta ................................... 75

5.2.2.

Limited data from persistence ............................................................. 76

5.2.3.

The case of the Birkirkara site .............................................................. 76

5.2.4.

Extent of tectonic affect....................................................................... 77

5.2.5.

Control on hydraulic conductivities from geologic contacts ............... 77

5.2.6.

Jointing link with nearest fault structure ............................................. 78

5.2.7.

Further data limitations ....................................................................... 78

5.3

Transmissivity .............................................................................................. 79

5.4

Potentiometry ............................................................................................. 80

5.5

Main Data Limitations ................................................................................. 80

5.6

Conceptual Ground Model .......................................................................... 81

Summary and Conclusions .................................................................................. 84


6.1

Main Conclusions ........................................................................................ 84

6.2

Further Studies ............................................................................................ 85

6.2.1.

Fault parameters and control .............................................................. 85

6.2.2.

Controls on jointing .............................................................................. 86

6.2.3.

Permeability variability of different formations and facies ................. 86

6.2.4.

Geochemistry ....................................................................................... 86

References.................................................................................................................. 87
Appendix A The Geological Map of Malta (1993) ................................................... 93
Appendix B Main water bodies as indicated by the Malta Resources Authority
(MRA) ......................................................................................................................... 95
Appendix C Mdina investigation boreholes from Gianfranco et al. (2003) .......... 105
Appendix D Stereonet plots .................................................................................. 116
Appendix E Full scan lines sheets .......................................................................... 126
Appendix F Tables and graphics of discontinuities aperture data ........................ 145
Appendix G Tables and graphs of discontinuities persistence data ..................... 162
Appendix H Relative hydraulic conductivities ....................................................... 172
Appendix I Hydrodynamic data ............................................................................. 174
Appendix J Fault data and structural contours of LCL .......................................... 179

Page | x

Tables
Table 1 Summary of the stratigraphical units of Malta (adopted from Pedley et al.,
1976; the Geological Map of Malta, 1993) ................................................................ 16
Table 2 Climate parameters monthly means (source: Sapiano et al., 2006) .......... 17
Table 3 Water balances for individual ground water bodies for the year 2003
(source: Sapiano et al., 2003). The groundwater body codes refer to groundwater
bodies as identified by the Malta Resources Authority, maps of which are presented
in Appendix B. ............................................................................................................ 18
Table 4 Average aquifer hydraulic properties compiled from literature ................ 23
Table 5 Table showing site names of sites included in study ................................. 26
Table 6 Summary of discontinuity data collected. Site reference numbers are
cross-referenced to Figure 11 and Table 5. Initials CS refer to Christian Schembri and
AM refer to Geotechnical Engineer Adrian Mifsud. .................................................. 27
Table 7 Summary of the main identifiable joint sets from pole concentrations for
each site and scan line. Site reference numbers are cross-referenced to Figure 11
and Table 5. ................................................................................................................ 28
Table 8 Summary of the main joint sets variability of averages ............................. 29
Table 9 Aperture ranges for each aperture width class. These aperture width
classes are used in the presentation of Appendices F and G graphs. These are the
basis for the analysis presented in Chapter 4. ........................................................... 30
Table 10 Persistence ranges for each persistence class. These persistence classes
are used in the presentation of Appendices F and G graphs. These are the basis for
the analysis presented in Chapter 4. ......................................................................... 30
Table 11 D and L co-ordinates used to determine a relationship in the form D=cL2
.................................................................................................................................... 34
Table 12 Expected maximum widths for a 36m displacement using predictive
equations from Michie et al. (2014) .......................................................................... 37

Page | xi

Figures
Figure 1 Location of the Maltese Islands is shown by the red circle (source: Google
Earth, 2014) .................................................................................................................. 1
Figure 2 Part of the North West coast of Malta as seen from Fomm ir-Rih Bay ... 2
Figure 3 Tectonic sketch of the Central Mediterranean. Approximate position of
Malta shown by the red circle. (source: Anzidei et al., 2001). Key details:
(1)Continental (a) and oceanic (b) parts of the Africa/Adriatic and Eurasian
forelands; (2) Tethyan belt comprised of oceanic remnants and intermediate
massifs (Pelagonian, Anatolian and Cyclades arcs); (3) deformation belts developed
on the African and Eurasian margins; (4) crustal thinning; (5) active thrust fronts; (6)
subduction zones; (7) inactive thrust fronts; (8,9,10) compressional, tensional and
transcurrent feautrues; (11) main trends of compressional deformation in the
Mediterranean Ridge and Calabrian Arc. (Anzidei et al., 2001)................................. 6
Figure 4 Isopach map of the LGL (source: Pedley et al., 1976) ................................. 7
Figure 5 Focal mechanisms with principal stress directions for various major fault
zone outcrops (source: Dart et al., 1993) .................................................................... 8
Figure 6 Diagram showing the North-South transfer fault zone between the
Pantelleria Trough to its west and the Malta Trough (also known as the Malta
Graben) and Linosa Trough to its east. These three structures form part of the
Pantelleria Rift (source: Argnani, 1990) ....................................................................... 9
Figure 7 Diagram showing the tectonic kinematics for the Hyblean-Malta Plate and
the Ionian Plate as proposed by Jongsma et al. (1987) (source: Jongsma et al., 1987)
.................................................................................................................................... 10
Figure 8 Simplified structural geologic map of the Maltese Islands (source: Dart et
al., 1993)..................................................................................................................... 13
Figure 9 Total annual rainfall for Luqa meteorological station (source: Sapiano et
al., 2006)..................................................................................................................... 17
Figure 10 Conceptual model of a solution subsidence structure (Pedley,
1975b:p.542) .............................................................................................................. 20
Figure 11 Sites visited indicated on the Geological Map of Malta (1993). Sites
marked in red include only a site reconnaissance exercise while magenta sites
include also the collection of discontinuity data. Discontinuity data for site 8 was
acquired and not carried out by myself. No site visit was paid to site 8. Site numbers
are cross-referenced to Table 5. ................................................................................ 25
Figure 12 Stereographic projection showing the entire joint set brackets
considered for grouping of all the discontinuity data overlaid over a scatter plot of
all the pole data. The inner circle represents a dip angle of 35o; poles within it
represent discontinuities with dip angles less than 35o. J9 & J10 are introduced to
include all of the data. ............................................................................................... 29

Page | xii

Figure 13 Points of which the details are presented in Table 11. The red line
indicates the direction along which the length of the fault is considered (adapted
from The Geological Map of Malta, 1993) ................................................................. 34
Figure 14 Graph showing growth path of a segmented isolated fault and the
complexity to determine this relationship due to field data scatter. For inter-linked
overlapping faults the difference between the green point and the blue point
depends on what fault length is considered. (adapted from Cartwright et al., 1995)
.................................................................................................................................... 35
Figure 15 Cumulative heave plot for a cliff scale section named Malta D with
location shown on map. Red lines on plot indicate single faults having major heave
spaced at approximately 600 metres. (adapted from Putz-Perrier & Sanderson, 2010)
.................................................................................................................................... 37
Figure 16 Aerial photo of Qammiegh Site (adapted from Google Earth, 2014). The
red polylines encircle slope instability areas which mainly involve rock toppling. The
blue polyline encircle a wetland and saltmarsh. Black lines show the locations of
faults and where they are dashed it means that they are inferred (The Geological
Map of Malta, 1993). Numbers are cross-referenced with photo numbers of this
section and the arrows show the orientation of view. .............................................. 38
Figure 17 Photo 1. Arrows indicate extent of main karst features and dashed lines
show enlarged joints observed from distance........................................................... 39
Figure 18 Photo 2 showing extensive cave development at Fault Zone. Dashed line
indicates inferred fault (The Geological Map of Malta, 1993). ................................. 40
Figure 19 Photo 3 showing UCL hanging wall ......................................................... 40
Figure 20 Photo 4 showing varying slope angles along a slope section. Red lines
indicate general slope angles of every formation. .................................................... 41
Figure 21 Aerial photo of L-Imgiebah Bay (adapted from Google Earth, 2014). The
red polylines encircle slope instability area. The green polylines encircle denser
vegetated areas. Numbers are cross-referenced with photo numbers of this section
and the arrows show the orientation of view. .......................................................... 42
Figure 22 Photo 1 showing cliff edge with the BC overlying the UGL. Dashed lines
indicate joints into the plane of the paper (approx. ENE-WSW), dotted polylines
indicate joints parallel to the plane of the paper (approx. NW-SE)........................... 43
Figure 23 Photo 2 showing desiccated clay surface............................................... 43
Figure 24 Photo 3 showing oxidised UGL joint as interpreted by Missenard et al.
(2014) ......................................................................................................................... 43
Figure 25 Photo 4 showing UCL cliff edge exhibiting at least two dominant
discontinuity sets. For explanation of annotations used refer to Figure 22. ............ 44
Figure 26 Aerial photo of Fomm ir-Rih Bay (adapted from Google Earth, 2014).
Numbers are cross-referenced with photo numbers of this section and the arrows
show the orientation of view. .................................................................................... 45

Page | xiii

Figure 27 Photo 1 showing Victoria fault at west coast outcrop. Annotations used
are explained by notes on photo. .............................................................................. 45
Figure 28 Photo 2 showing detail of antithetic fault shown in Figure 27. ............. 47
Figure 29 Photo 3 showing LGL outcrop at location of Scan line A (scale shown
by geologic hammer) .................................................................................................. 47
Figure 30 Photo 4 shows MGL at location of scan line B. Dashed lines show
examples of sub-horizontal desiccation joints (scale shown by field notebook) ...... 47
Figure 31 Photo 5 showing the Lower Main Phosphorite Conglomerate (scale
shown by field notebook scale line) .......................................................................... 47
Figure 32 Karst caves in the Attard member of the LCL at Wied il-Ghasel where it
cross-cuts the Victoria fault, the rock face strike is approximately NNE................... 48
Figure 33 Aerial photo of site visited at Gharghur (adapted from Google Earth,
2014). The green line shows a stretch of denser vegetation. Numbers are crossreferenced with photo numbers of this section and the arrows show the orientation
of view. ....................................................................................................................... 49
Figure 34 Photo 1 showing formed notches at Xlendi member of the LCL. Dashed
polyline indicates a wider aperture joint. .................................................................. 50
Figure 35 Photo 2 showing detail of calcite deposition .......................................... 50
Figure 36 Photo 3 showing karstified cave pillar .................................................... 50
Figure 37 Infilled joint having wide aperture .......................................................... 51
Figure 38 Fault Breccia ............................................................................................ 51
Figure 39 Wide open aperture at fault zone ........................................................... 51
Figure 40 Schematic plan layout of Excavation site visited at Msida with an
indication of the excavation faces references .......................................................... 53
Figure 41 Minor Karst feature noted at wall A ....................................................... 53
Figure 42 Excavation face A with dotted lines indicating discontinuity zones (scale
is shown by excavator) ............................................................................................... 53
Figure 43 Excavation face B with dotted lines indicating discontinuity zones (scale
is shown by excavator) ............................................................................................... 54
Figure 44 Excavation face D on the left side, face B on the right side and face C in
between. Discontinuity zones are indicated by dotted lines (scale is shown by
excavator)................................................................................................................... 54
Figure 45 Aerial photo of site visited at Xghajra (adapted from Google Earth, 2014).
Dashed lines show examples of NE discontinuities which can be observed even from
this aerial photo. ........................................................................................................ 55
Figure 46 Photo 1 showing NE trending discontinuities on the LGL wall and on the
LCL ground (scale shown by field notebook. Dashed lines show discontinuity dip
angles. The polyline shows a karstified stretch on the LCL ground. .......................... 56

Page | xiv

Figure 47 Photo 2 showing contact between LGL and LCL and Scutella echinoids
marker (scale shown by field notebook).................................................................... 56
Figure 48 Aerial photo of Munxar Site (adapted from Google Earth, 2014) .......... 57
Figure 49 Photo 2. Dashed lines show examples of tight discontinuities which may
be interpreted as sedimentation desiccation discontinuities. (scale shown by field
notebook scale line) ................................................................................................... 57
Figure 50 Photo 1 showing a low MGL cliff face ..................................................... 58
Figure 51 Investigation boreholes location encircled in red (source: Google Earth,
2014) .......................................................................................................................... 58
Figure 52 MGL outcrop at Fomm ir-Rih Bay scan line B. Dashed lines show main
joints. .......................................................................................................................... 60
Figure 53 Cave-like structure adjacent to the position of the start of scan line A . 62
Figure 54 Photo of Birkirkara site showing undulating discontinuity (scale shown
by mobile crane) ........................................................................................................ 63
Figure 55 Variability of transmissivity with borehole depth below top of LCL (line
shown is the trend line) ............................................................................................. 67
Figure 56 Variability of transmissivity with borehole depth with respect to the
mean sea level (line shown is the trend line) ............................................................ 68
Figure 57 Variability of transmissivity in relation to distance away from fault
considered at surface. Black line shows a possible trend line if data points above the
red arrow and in the circle are ignored. .................................................................... 69
Figure 58 Variability of transmissivity in relation to distance away from fault
considered at depth of borehole end. Black line shows a possible trend line if data
points above the red arrow and in the circle are ignored. ........................................ 69
Figure 59 1990 potentiometric map superimposed on the Geological Map of Malta
(1993). Dashed lines show main faults average alignments, dots with number show
locations of gauged boreholes with water piezometric level. (adapted from BRGM,
1991c & the Geological Map of Malta, 1993) ............................................................ 71
Figure 60 The Victoria fault within the regional geology (adapted from the
Geological Map of Malta, 1993)................................................................................. 78
Figure 61 Red border shows the area of the conceptual ground model presented
in this section (adapted from the Geological Map of Malta, 1993) .......................... 81
Figure 62 Conceptual Ground Model highlighting regional hydrogeology of Malta.
Annotations cross-referenced to numbers are included on the next page. ................ 82

Page | xv

Introduction

The Maltese Islands are located in the Mediterranean Sea, about 90km south of
Sicily and about 300km east of Tunisia. The location of the Maltese archipelago is
shown (Figure 1).

250km

Figure 1 Location of the Maltese Islands is shown by the red circle (source: Google Earth, 2014)

The geologic strata consist in marine sedimentary and are predominantly made up
of carbonates. Five main stratigraphic units are identified with the oldest being the
Lower Coralline Limestone (LCL) of Chattian age from the Oligocene and the
youngest the Upper Coralline Limestone (UCL) from the early Messinian period of
the Miocene. The Coralline Limestones usually form bare karstic plateaux in the
landscape while the Globigerina Limestone (GL) produces gentler landscape (Pedley
et al., 1976). In exposed Blue Clay (BC) slopes, drainage gullies and rock toppling of
the overlying UCL can be observed especially along the North West coast of Malta
(Devoto et al., 2012; Figure 2). The BC is the most fertile unit (Pedley et al., 1976)
owing in part to its low permeability characteristics. The soil produced from the
water solution of the Upper Coralline Limestones tends to be fertile too (BRGM,

Page | 1

1991) and that explains the fact that a good number of solution subsidence
structures are occupied by agricultural fields.

60m

Figure 2 Part of the North West coast of Malta as seen from Fomm ir-Rih Bay

North of the Victoria Fault, Malta is dominated by horst and graben structures. Two
main fault sets outcrop on Malta with one trending approximately ENE-WSW and
the other NW-SE.
Malta has two main aquifer types. The upper perched aquifer overlying the BC
formation which acts as an aquitard and the lower being the mean sea level aquifer
of the Ghyben-Herzberg Lens type (Alexander, 1988; ATIGA, 1972; BRGM, 1991c;
Sapiano et al., 2006; Stuart et al., 2010). The BC formation can be subject to some
seepage losses which provide for some degree of connectivity between the two
aquifers. The upper perched aquifer is discontinuous and is made from several
blocks (Appendix B).
In several parts of the literature it is highlighted that the aquifers are dual porosity
with the primary porosity referring to the matrix permeability while the secondary
refers to the permeability due to the fractures which may be altered by karstic
carbonate dissolution (Newbery, 1968; BRGM, 1991; Sapiano et al., 2006;
Bakalowicz & Mangion, 2003; Stuart et al., 2010).

Page | 2

This dissertation has the aim of developing a good understanding of structural


controls on the hydrogeology of Malta as the main island of the Maltese Islands.

1.1

Scope of Work

The importance of water to support life is undebatable. During the last decades
Malta has witnessed increasing demand for water resources while water recharge
remained low. The unbalance between water demand and supply has sometimes
resulted in diminishing water quality. It is therefore to no ones surprise that
hydrogeologic studies of Malta generally deal with the hydrologic balance, water
quality and water management strategies (ATIGA, 1972; BRGM, 1991; Sapiano et al.,
2006; Stuart et al., 2010). The study of structural controls on the hydrogeology of
Malta has therefore been a neglected subject. However a new interest from the
petroleum industry to characterise the hydraulic properties of Maltas geology
seems to be on the rise. Missenard et al. (2014) describe Malta as an open
laboratory of the Mediterranean which can possibly provide hints for offshore
explorations in the Mediterranean region.
Detailed studies from the petroleum industry even though their interest is not
geared towards understanding the hydrogeology may provide good data that can
be used to better understand the hydrogeology. In order to better understand the
structural controls on the hydrogeology of Malta an extensive literature review and
desk study coupled by a limited amount of field work are carried out. This scope
includes:

the identification of main controls from geomorphologic site reconnaissance,

a study of scan line discontinuity data carried out at random sites with an
analysis of the main parameters, how these relate to the regional geologic
environment is speculated and an indication of the variation of relative
hydraulic conductivity is given,

Page | 3

re-interpretation of some spatial hydrodynamic and potentiometric data,


and

an interpretation of the main fault controls on the hydrogeology of Malta by


combining field data with previous studies presented in a regional
conceptual ground model.

1.2

Overview of Work

The first chapter of this work gives a general background to the geology of Malta,
outlines the main aim and gives an overall overview of the work.
The second chapter presents a literature review giving a sound geologic and
hydrogeologic background of Malta. The geologic background includes tectonic
setting, tectonic history, structural geology, sedimentation environment and a brief
description of the main geologic strata. The hydrogeologic background includes
some basic hydroclimatological and water balance data, hydrogeologic setting
including the main aquifer types and some geomorphologic features realted to
hydrogeology, basic hydrogeologic characteristics of the different geologic materials,
basic hydraulic aquifer data, some observations from fault effect and from
geochemical studies.
The third chapter explains the main techniques used for data collection and its
analysis. It highlights how the methods are applied, developed and what their
limitations may be. This chapter also includes explanation of how data is going to be
presented in chapter 4. The development of a method to create a framework to
predict fracture layouts along the Victoria fault is attempted. This part of the
process serves to highlight the complexity involved and to highlight data gaps,
rather than serving the purpose of successfully providing a detailed method that
can be applied to predict fracture layouts. This part is still deemed to be very useful
in understanding the realm of structural control on the hydrogeology of Malta
especially in the piecing together of a regional conceptual ground model.
Page | 4

The fourth chapter presents observations and the analysis carried out. The topics
include geomorphologic site reconnaissance, analysis of spatial discontinuity data, a
spatial analysis of transmissivity and a small note on the potentiometric data
available. During the course of this chapter observations from one site may crossreference observations from another site or from literature. In so doing the pace is
set for the fifth chapter.
The fifth chapter provides the main discussion with all the observations brought
together providing a more holistic understanding of the hydrogeology of Malta. A
regional conceptual ground model is presented with this scope. The uncertainties,
data limitations, further implications and some suggestions for further studies are
included along the discussion.
The final chapter presents a summary of the main conclusions and
recommendations for further study.

Page | 5

2
2.1

Geologic and Hydrogeologic Background of Malta


Geologic Background
2.1.1.

Tectonic Setting

500km

Pantelleria Rift

Figure 3 Tectonic sketch of the Central Mediterranean. Approximate position of Malta shown by the red circle.
(source: Anzidei et al., 2001). Key details: (1)Continental (a) and oceanic (b) parts of the Africa/Adriatic and
Eurasian forelands; (2) Tethyan belt comprised of oceanic remnants and intermediate massifs (Pelagonian,
Anatolian and Cyclades arcs); (3) deformation belts developed on the African and Eurasian margins; (4) crustal
thinning; (5) active thrust fronts; (6) subduction zones; (7) inactive thrust fronts; (8,9,10) compressional,
tensional and transcurrent feautrues; (11) main trends of compressional deformation in the Mediterranean Ridge
and Calabrian Arc. (Anzidei et al., 2001)

The Maltese Islands form part of the African continental plate where MioceneQuaternary extensional basins system formed to accommodate extension in the
foreland of the Apennine-Maghrebian thrust and fold belt (Dart et al., 1993; Argnani,
1990). The latter marks the collision zone between the African and Eurasian plates
to the north of the Maltese Islands. A tectonic sketch of the central Mediterranean
region is presented (Figure 3 sourced from Anzidei et al., 2001). This diagram gives a
general idea of the main geological structures of the region however the individual
details are the subject of various debates. At a micro plate scale the Maltese

Page | 6

archipelago forms part of the Hyblean-Malta Plateau (marked IB on Figure 3; Pedley


et al., 1976).
An isopach map of the Lower Globigerina Limestone (LGL) is shown in Figure 4
(Pedley et al., 1976). This map shows thicker sections of the LGL member, which is
Aquitanian in age, close to the centre of Malta. This occurrence can be attributed to
up-arching of this member (Pedley, 1987) due to a compression action preceding
the rifting with the latter event estimated to start at approximately 21 million years
ago (Dart et al., 1993). Although this statement is plausible one would desire more
evidence to support it, however it is felt that this is outside the scope of the present
study.

Figure 4 Isopach map of the LGL (source: Pedley et al., 1976)

A detailed study including both onshore and offshore data proposes a tectonic
hypothesis involving a north-south rift direction as responsible for the two main
fault sets found on the Maltese Islands (Dart et al., 1993). Each fault set is made up
Page | 7

of two fault sub-sets, each with the same approximate strike but with opposite dip
direction forming bounding edges of basins. They argue that there are two possible
scenarios where a single direction of stress can be responsible for this occurrence. It
can be either the re-activation of pre-existing discontinuities or a state of tri-axial
strain with the minor strain not equal to zero. In support of this hypothesis they
collect discontinuity scan line data including also slip characteristics at major fault
zones. They followed the kinematical methods proposed by Marrett & Allmendinger
(1990) and obtained focal mechanisms with the extensional stress direction being
approximately North-South for all fault exposures (Figure 5 sourced from Dart et al.,
1993).

Figure 5 Focal mechanisms with principal stress directions for various major fault zone outcrops (source: Dart
et al., 1993)

This interpretation conforms to the hypothesis of Argnani (1990), who suggests a


North-South transfer fault zone located within the Pantelleria Rift (Figure 6 sourced
from Argnani, 1990). This transfer zone accommodates differential North-South
extension and its interpretation is supported by the occurrence of strike-slip
indications along it, such as volcanic centres and a positive flower structure at the
north limb of the Pantelleria Rift. However their interpretation may contrast with
interpretations by other authors for other structures in the region.
Page | 8

Figure 6 Diagram showing the North-South transfer fault zone between the Pantelleria Trough to its west and
the Malta Trough (also known as the Malta Graben) and Linosa Trough to its east. These three structures form
part of the Pantelleria Rift (source: Argnani, 1990)

Another interpretation for the kinematics of the area is given by Jongsma et al.
(1987). The authors attribute the opening of the Pantelleria Rift (also known as the
Medina Wrench zone) as a pull-apart basin due to dextral strike-slip movements.
This structure accommodates the faster movement of the Hyblean-Malta Plate and
the Ionian Plate towards the east when compared to the African and Eurasian Plates.
A component of anti-clockwise rotation about poles in the South of Italy is also
reported (Figure 7 sourced from Jongsma et al., 1987).
Grasso & Reuther (1988) as cited by Dart et al. (1993) propose that the Pantelleria
Rift formed as a pull-apart basin to accommodate strike-slip motion along the NNESSW trending Scicli fault to the North of Malta and towards the South East corner of
Sicily.

Page | 9

Figure 7 Diagram showing the tectonic kinematics for the Hyblean-Malta Plate and the Ionian Plate as
proposed by Jongsma et al. (1987) (source: Jongsma et al., 1987)

The latter two hypotheses are not supported by much geomorphologic expression
neither in the Maltese Islands nor within the Pantelleria Rift (Dart et al., 1993;
Argnani, 1990). Minor exposures of strike-slip motion are witnessed in Gozo
including small scale strike-slip damage zones from the north west of Gozo (Kim et
al., 2003) and strike-slip horsetail faulting near Qala with displacements less than 1
centimetre (Illies, 1981; Pedley et al., 1976). Some features such as solution
subsidence structures near Dwejra in Gozo were interpreted as strike-slip
manifestations (Illies, 1981). A contrasting interpretation is that these faults were
formed as a cause of two solution subsidence structures in the area (Pedley, 1975b).
From extensive field studies the faults trending approximately WNW-ESE are

Page | 10

deduced to be predominantly normal dip-slip faults (Putz-Perrier, 2008; Putz-Perrier


& Sanderson, 2010; Michie et al., 2014).
Gardiner et al. (1995) reports Upper Pliocene right-lateral transtension at the North
of Gozo graben with right-stepping ridges. This activity probably reactivated the
older normal fault sets. The tectonic structural block model proposed for the
Hyblean-Malta plateau by Gardiner et al. (1995) could well combine both the
interpretations of Dart et al. (1993) and Jongsma et al. (1987) with possibly having a
rotation of the extensional axis from north-south towards north east-south west in
the last 5 million years with minimal evidence on the Maltese Islands.
The uplift of the Maltese Islands occurred from the late Messinian to the midPliocene. It is believed to be linked with the reactivation of the ENE-WSW faults by
right-lateral wrenching coupled with sea lowering during the Messinian (Pedley,
1987; Pedley, 2011). The stratigraphic units of Malta have a sub-horizontal bedding
dip towards the north-east in various areas (Pedley et al., 1976; Pedley, 1987; Dart
et al., 1993). This can also be indicative of a rising of the west coast of Malta and a
drowning of the east coast. This is strengthened by an observation of stalagmites at
the sea bottom noted during construction of the Valletta breakwater by Rizzo (1932)
as cited by Trechmann (1938). Stalagmites do not form under water.

2.1.2.

Tectonic History Debates

There are a number of historical interpretations of how the two fault families
outcropping in Malta. Just a small reminder, the two main fault families are the
ENE-WSW and the NW-SE trending faults.
Gardiner et al. (1995) suggest that the NW-trending Malta Trough forms first as a
reaction to relieve tensional stress during the collision of the African Plate with the
Eurasian Plate. The ENE-WSW trending faults of the Maltese region are attributed
to a mid-Pliocene regional uplift from Gozo to SE Sicily. With their interpretation

Page | 11

these authors believe that the NW-SE faults formed first followed by the ENE-WSW
faults.
Other parts of the literature suggest two rift systems with the first producing the
ENE-WSW faults and the second forming the NW-SE fault trends (Illies, 1981;
Reuther & Eisbacher, 1985).
Dart et al. (1993), on the other hand, report that both fault sets formed
contemporaneously supporting their statement with field data. These authors
noted instances of both fault sets cross-cutting each other, single striae lineation
per fault and similar depositional patterns in both the North Gozo Graben (ENEWSW oriented basin) and the Pantelleria Rift (NW-SE oriented basin).
The extensional rifting is estimated to start at approximately 21 Million years ago
with the major extensional rifting probably taking place between 5 Million years ago
and 1.5 Million years ago (Dart et al., 1993). When rifting ceased some parts of the
literature suggest dextral strike-slip re-activation of the ENE-WSW faults due to the
rotation of the extensional stress axis more towards the north-east due to further
continental plates collision (Illies, 1981; Reuther & Eisbacher, 1985; Gardiner et al.,
1995). Even though there is minimal evidence of strike-slip faulting onshore Malta,
the latter statement should not be ignored.

2.1.3.

Onshore Structural Geology of Malta

From the Geological Map of Malta (1993), which is attached in Appendix A, one can
observe that the archipelago has two main sets of faults.
One set strikes approximately between N050o and N090o and is mostly evident over
a 14 km stretch between the Victoria fault (VLF) and the South of Gozo Fault (SGF or
Qala Fault). This area is dominated by successive horst and graben structures which
together form the North Malta Graben. The largest throws of this set are reported
at the Victoria fault and are approximately 195 to 200 metres at the west coast fault

Page | 12

zone and at the east coast the total displacements of the fault zone are about 90
metres with 60 metres displacements occurring on the Victoria Fault alone (Pedley
et al., 1976; Costain, 1957-1958 as cited by Dart et al., 1993; Reuther & Eisbacher,
1985; Michie et al., 2014). It is also reported that faults are not identified by
offshore seismic sections to the east of Malta therefore we expect throws to be less
than 10 metres in this region however we do not know where these seismic
sections were carried out (Dart et al., 1993).
The other set strikes approximately between N120o and N140o with its outcrops
being rare with one excpetionally good outcrop at Il-Maghlaq Fault (IMF) to the
south west of Malta (Michie et al., 2014; Reuther & Eisbacher, 1985). This set trends
sub-parallel to the Pantelleria Rift. Il-Maghlaq Fault is reported to have the highest
vertical throws observable in the Maltese Islands with over 210m (Reuther &
Eisbacher, 1985; Bonson et al., 2007). A simplified structural geologic map is
presented (Figure 8 sourced from Dart et al., 1993).

Figure 8 Simplified structural geologic map of the Maltese Islands (source: Dart et al., 1993)

From field observations carried out in Malta it was noted that fault zone widths are
narrow in plan and that they contain both synthetic and antithetic faults (Dart et al.,
1993). Predictive equations for fault zone widths in relation to fault displacements
are proposed (Michie et al., 2014). Smearing of the Blue Clay is reported in faults

Page | 13

with throws of approximately greater than fifty metres, while for intermediate
throws a wider zone of deformation within the Blue Clay with brittle structures is
reported (Missenard et al., 2014). Some minor synclines are also noted close to
major faults and their occurrence is attributed to fault drag (Pedley et al., 1976).
From seismic sections it was deduced that maximum throws in the North Gozo
Graben and the Pantelleria Rift are estimated at 1600m and 2200m respectively
(Dart et al., 1993). This shows us higher offshore activity in the mentioned regions
than onshore Malta. This is also shown from the calculated regional strains. Average
regional extensional strains over the entire Maltese Islands are deemed to be
approximately 3% strain (Putz-Perrier, 2008; Putz-Perrier & Sanderson, 2010), while
extensional strains of about 10% and 17% were reported for the North Gozo Graben
and the Pantelleria Rift respectively (Dart et al., 1993).

2.1.4.

Main Stratigraphical Units

The geologic formations of Malta consist mostly of sedimentary marine carbonates


deposited at shallow sea depths with the highest sea depths estimated not to be
greater than 250 metres (Pedley et al., 1976; Bonson et al., 2007). A summary of the
stratigraphical succession of Malta is presented (Table 1 sourced from Pedley et al.,
1976 and the Geologic Map of Malta, 1993). A few mineralogical tests have shown
that even the Blue Clay Formation has about 15-25% of calcium carbonates
(Missenard et al., 2014). Calcium carbonates if subjected to solution by water could
compromise the seal provided by BC (BRGM, 1991c). Maltese carbonates are very
similar to carbonates found in Sicily and in the Sirte Basin of Libya (Pedley et al.,
1976). Information on the pre-Miocene strata is only limited to an exploration
borehole at Naxxar (Pedley et al., 1976; Dart et al., 1993).
Pliocene strata are extensively thick in the Pantelleria Rift and North Gozo Graben
while they are absent from Malta. This shows the large difference in elevation

Page | 14

between these areas during the uplift of Malta above sea level (Dart et al., 1993;
Trechmann, 1938).
The post-Miocene strata of Malta include mammal remains which are comparable
with deposits in SE Sicily which has undergone similar terrestrial processes (Pedley,
2011). These deposits are quite discontinuously distributed and were thus
subsequently neglected by many studies. A marker bedding was identified as the
San Leonardo Marine Abrasion Surface, which was deemed to provide a good
starting point to solve the recent geological history of Malta.
Members and facies are distinguished within the geologic formations. Facies occur
due to varying sea depths of what has been idealised as a ramp profile with subhorizontal dip angles (Pedley, 1998). This author highlights that facies variability
such as grain size and types of marine deposits relate to depth of deposition and
direction of predominant sea current. Larger grain sizes are expected at shallower
parts of the ramp, while marls and finer grains are expected at the outer part of the
ramp. Marls would also be expected at the shallowest parts of the ramp if the ramp
is facing sea currents.
Effects of tectonics are also noted from palaeolandslides when the sediments were
still in a semi-lithified state (Pedley, 1998). From a study of facies in the UCL it is
believed that the palaeoenvironments are mostly controlled by faulting followed by
sea currents (Bosence & Pedley, 1982).
Within the GL two phosphorite conglomerates and hardgrounds are reported and
studied (Pedley & Bennett, 1985; Pratt, 1990). Both conglomerates include clasts
from hardground material with the source area identified to be to the west and
north of modern Malta (Pedley & Bennett, 1985). The bottom conglomerate lies
directly over a hardground (Pedley & Bennett, 1985). Hardgrounds occurrence is
attributed to an increase in sea energy due to a sea-level drop which disturbed fine
material while this was coupled by a lack of deposition (Pratt, 1990).

Page | 15

Mtarfa
(12-16m)

75m)

(15-

(BC)

Ghajn Melel
(0-13m)

Clay

Upper Coralline Limestone (UCL)

Formation
& Member
Imbark
(4-25m)
Tal-Pitkal
(30-50m)

Globigerina Limestone (GL)

Upper (U)
(8-26m)

Lower Coralline Limestone (LCL)

Chattian

Oligocene

Aquitanian

Burdigalian

Langhian

Blue

Miocene

Tortonian

Early
Messinian

Geological Age

Middle (M)
(15-38m)

Lower (L)
(0-80m)

Description
Hard, pale-grey carbonates with sparse faunas.
Pale grey and brownish-grey wackestones and
packstones. Contains coralline algal, mollusc, echinoid
bioclasts and rhodoliths. Upper beds dominated by
carbonate mudstones.
Massive to thickly bedded carbonate mudstones and
wackestones. Unconformable upon Greensand in
western outcrops. Carbonates become white and chalky
in the upper two thirds of eastern outcrops. Contains
rhodoliths.
Massive bedded dark to pale brown foraminiferal
packstones. Contains glauconite above a basal UCL
erosion surface in West Malta. The glauconite-rich
Greensand Formation is included in this member for
convenience as it rarely exceeds 1 metre thickness.
Medium grey pelagic marls, typically with pale bands rich
in planktonic foraminifera but lower clay content.
A fine grained planktonic foraminiferal limestone
sequence made up of a central pale grey marl layer, a
lower and an upper cream coloured wackestone. A
phosphorite conglomerate bed occurs at the base. Lies
conformable in eastern outcrops but lies above a
hardground and erosion surface in the west.
A planktonic foraminifera-rich sequence of massive,
white, soft carbonate mudstones locally passing into
pale-grey marly mudstones. Base is unconformable over
lower GL member.
Pale cream to yellow planktonic foraminiferal packstones
becoming wackestones above the base which is
phosphatised in the west and includes a conglomerate
bed. The top of the member is marked by a hardground.

Il-Mara
(0-20m)

Tabular beds of pale-cream to pale-grey carbonate


mudstones, wackestones and packstones. The top of the
member is transitional with the LGL. Bryozoan (moss
animals) fragments are common.

Xlendi
(0-22m)
Attard
(10-15m)

Planar to cross-stratified, coarse-grained packstones with


abundant coralline algal fragments.
Grey wackestones and packestones. Large coralline algal
rhodoliths are widespread. An extensive N-S trending belt
of patch-reefs extend from Wied Maghlaq to Naxxar.
Massive bedded, pale yellowish-grey carbonate
mudstones are dominant and foraminifera are frequent.
It passes transitionally up into the Attard Member.

Maghlaq
(>38m)

Table 1 Summary of the stratigraphical units of Malta (adopted from Pedley et al., 1976; the Geological Map
of Malta, 1993)

Page | 16

2.2

Hydrogeologic Background
2.2.1.

Basic Hydroclimatological data

The climate is semi-arid Mediterranean with summers hot and dry while winters
mild and wet (Sapiano et al., 2006). The annual rainfall totals for the Luqa
meteorological station covering the period 1947-2004 (Figure 9 sourced from
Sapiano et al., 2006) and the average monthly rainfall for all of Malta (Table 2
sourced from Sapiano et al., 2006) are shown.

Figure 9 Total annual rainfall for Luqa meteorological station (source: Sapiano et al., 2006)

Table 2 Climate parameters monthly means (source: Sapiano et al., 2006)

Page | 17

Even though the average annual rainfall is not high, flooding of a good number of
main streets to the east of Malta, especially after the post-summer flash storms, has
been an issue for several years to such an extent that a National Flood Relief Project
was put forward (Ministry for European Affairs, 2014). Older systems to tackle this
problem included a number of small scale dams along the water courses to store
rainwater run-off while at the same time encouraging infiltration (Sapiano et al.,
2006).

2.2.2.

Water Balance

Hydrological balance estimates for each individual aquifer for the year 2003 show
an unsustainable state of the major aquifers that are the mean sea level aquifers
(Sapiano et al., 2003).

Table 3 Water balances for individual ground water bodies for the year 2003 (source: Sapiano et al., 2003).
The groundwater body codes refer to groundwater bodies as identified by the Malta Resources Authority, maps of
which are presented in Appendix B.

2.2.3.

Hydrogeological Setting

As previously highlighted Malta has two main types of aquifers. The perched aquifer
which overlies the BC formation is generally very thin with flow towards the down
Page | 18

dip of the same stratum (Newbery, 1968). Springs are noted at the UCL and BC
interface. Variable mechanical and hydrogeological properties of the brittle UCL and
the more plastic underlying BC has led to slope instabilities such as rock toppling at
the UCL peripheries (Gianfranco et al., 2003; Magri et al., 2008; Devoto et al., 2012).
The mean sea level aquifer consists of a lens of freshwater which floats over sea
saltwater with the main host rock being the LCL. The UCL may also be a host rock of
this aquifer when it lies at the sea level in the horst and graben structure of north
Malta (Alexander, 1988). The main difficulties for this aquifer to meet the demands
of the potable water supply include the relatively small recharge area and extents of
the aquifer itself, the large population density, seawater intrusions due to over and
uncontrolled pumping and several nitrate sources of pollution (Sapiano et al., 2006;
Stuart et al., 2010).
Surface drainage channels are noted to follow Maltas fault trends and fracturing is
also noted to occur approximately parallel to the main ENE-WSW faults (Alexander,
1988; Gutierrez, 1994 as cited by Bakalowicz & Mangion, 2003). It has been noted
that surface water drains rapidly mostly through karst rock features (Sapiano et al.,
2006; Stuart et al., 2010). Contrastingly it has been also noted that karst plays a
minimal role in the aquifer recharge (Bakalowicz & Mangion, 2003). These two
statements highlight the difficulty to generalize and quantify the effect karst
features may have.
Major surface karst features involve circular subsidence structures (Trechmann,
1938; Newbery, 1968; Newbery 1975; Pedley, 1975b; Alexander, 1988). Two
formation methods of such structures are proposed. The first mode involves a
cavern roof collapse subsiding the overlying material. Caverns eventually enlarge
the process of which would be accelerated by faulting and subsequent subsiding
would occur (Newbery, 1976; Pedley, 1975b). The second mode may involve the
softening of the BC at larger inflow zones from the overlying UCL (Newbery, 1976).
A conceptual model as proposed by Pedley (1975) is shown (Figure 10). It is not easy
to say whether these structures are fault controlled from surface geomorphologic

Page | 19

observations; however this may be more obvious from observations in underground


galleries (Newbery, 1976).

Figure 10 Conceptual model of a solution subsidence structure (Pedley, 1975b:p.542)

It has been noted that different stratigraphic units may have different karst
development potentials with the fine-grained materials such as the GL having
localised vertical enlargement of fractures but with limited horizontal extent while
wider zones in all directions may develop within the Coralline Limestones (Newbery,
1968; Pedley, 1975b; Bakalowicz & Mangion, 2003; Stuart et al., 2010). Generally
speaking, in mean sea level aquifers one may also expect karst features at the
saline-freshwater contact zone (Mylroie & Mylroie, 2007).

2.2.4.

Effects of faults on fluid flow

The BC formation is considered to form an impermeable layer (Stuart et al., 2010)


however some groundwater from the perched aquifers has been deemed to flow to
the mean sea level aquifers through fracture zones (Sapiano et al., 2006). Flow
through the BC formation can also occur as seepage facilitated by the formations

Page | 20

carbonate content and solution subsidence structures (BRGM, 1991c; Stuart et al.,
2010).
Missenard et al. (2014) studied palaeo-fluid circulations. The authors made
distinctions between faults with throws less than 5 metres (low throw), faults with
throws in the range of 5 to 50 metres (medium throw) and faults with throws
greater than 50 metres (large throw). No fluid flow is observed at the low throw
faults, which observation suggests the ability of the BC stratum to stop the faults
from propagating through its thickness in such a case and hence retain its sealing
ability. Likewise no fluid flow is observed at the large throw faults which
observation suggests that palaeo-fluid flow started at around the Late Miocene
triggered by the uplift and sea level drop of the Messinian event while these faults
were already sealed by the clay smears. However palaeo-fluid flow through the
medium throw faults was observed in the BC in the form of gypsum filled
discontinuities and oxidised bands of clay surrounding these zones. In this case a
breach of the sealing capacities of the BC is suggested.
At South of Gozo fault (or Qala fault), the BC seal was observed to be breached such
that the UCL was adjacent to the LCL (Newbery, 1968). Reported aquifer thicknesses
in proximity to this fault were at a maximum of about 40 metres above the BC
(Newbery, 1968). This observation gives scope to carry out studies such as that
carried out by Micarelli et al. (2006) at the South East of Sicily which investigated
permeability reduction at different distances from the fault core with both hanging
wall and footwall made up of carbonate limestones. The responsible mechanisms
for this occurrence as highlighted by Micarelli et al. (2006) could include pore
collapse, grain crushing, rotation-enhanced abrasion and calcite precipitation.
Faults may have dual functions, they may act as a seal in the fault core formed or
they may act as flow conduits in the damaged zones. Seebeck et al. (2014) studied
the relationships between proximity to fault, fracture density and permeability in
sandstone formations. The concept of a critical fracture density at which spikes of
an increased permeability occur is discussed.

Page | 21

The topic of this section is extensively studied in the literature and this makes it
clear that the faults function with respect to fluid flow is not straight forward to
determine and requires good studies of the fault architectures followed by more
specific field and lab testing (Sapiano et al., 2006; Bonson et al., 2007; Michie et al.,
2014). Bonson et al. (2007) and Michie et al. (2014) have done extensive mapping of
the two main faults found in Malta, being Il-Maghlaq fault and the Victoria fault
respectively. Their data and observations would be used in order to propose a fault
architecture model along the Victoria fault.
Previous documents on the hydrogeology of Malta have said that the Victoria fault
does not form a sealed boundary except at some locations which are expected to
be near the west coast end. Similarly Il-Maghlaq Fault forms an impermeable seal at
only some locations (BRGM, 1991c; Sapiano et al., 2006).

2.2.5.

Aquifer hydraulic properties

Water flow through the GL is mostly controlled by fractures and this formation
provides very scarce locations that are fractured enough for possible water
production (Sapiano et al., 2006; Stuart et al., 2010). This formation has high
porosity but low permeability. The GL may provide a degree of confinement to the
underlying mean sea level aquifer depending on its bottom level with relation to the
sea level (Stuart et al., 2010). The MGL being a marly limestone probably has the
lowest permeability of the formation (Sapiano et al., 2006; Stuart et al., 2010).
Facies of the LCL having a higher concentration of coral-reef formations are
generally more porous and permeable however when compared with the
characteristic coral reefs their permeability is lower (Sapiano et al., 2006;
Bakalowicz & Mangion, 2003). Transmissivity can reach values of 1000m2/d for the
LCL where discontinuities may have been subject to dissolution (BRGM, 1991b;
Stuart et al., 2010). Well test data presented in BRGM (1991b) was used for part of
the analysis of this study. Prior to this data only two well pumping tests were

Page | 22

carried out (ATIGA, 1972). A compilation of average aquifer hydraulic properties are
presented in Table 4 (BRGM, 1991b; Bakalowicz & Mangion, 2003; Stuart et al.,
2010). This shows the lack of detail available in this regard even more so when
generally not much detail is given about what type of test is carried out and no
geologic description.
Parameter
Primary Porosity (%)
Effective Porosity (%)
Intact rock permeability (m/s)
Aquifer hydraulic permeability (m/s)
Transmissivity (m2/s)

UCL
41-45

GL
32-40

5.9x10-7

1.5x10-7

LCL
7-20
10-15
4.05x10-4
0.1 2x10-5

Table 4 Average aquifer hydraulic properties compiled from literature

Since the geologic strata were deposited at relatively shallow depths, pressure
solution does not occur (Bonson et al., 2007).

2.2.6.

Geochemical Studies

Stuart et al. (2010) carried the most recent geochemical study of Maltas
groundwater. From testing for the waters age it was discovered that the perched
aquifer has a faster response to rainfall with a mean saturated age of fifteen years.
The mean sea level aquifer has a mix of older and younger waters with the means
being fifteen and forty years. This is indicative of having two concurrent flow
mechanisms in the latter aquifer, one being slow through the rock matrix and the
other a rapid flow through discontinuities.
Similar conclusions were reached by a previous study from chemical results at
Fiddien borehole which is located at the Rabat-Dingli plateau which resulted in the
presence of modern desalinated seawater at both aquifers (Bakalowicz & Mangion,
2003). Higher concentrations were noted at the mean sea level aquifer with a
reduction in the thickness of the unsaturated zone (Stuart et al., 2010). From this
data it is plausible to conclude that the groundwater flow cycle depends on the
state and thickness of impermeable cover (Stuart et al., 2010). This observation is

Page | 23

also consistent with data from Gozo which has wider BC and MGL cover and
consequently higher mean saturated ages between thirty and sixty years (Stuart et
al., 2010). For Malta, the geographical centre was noted to have the oldest relative
age (Stuart et al., 2010).
It is plausible to believe that fast rainfall response in the perched aquifer is
indicative of higher transmissivities even though no field testing has been carried
out in this regard (Stuart et al., 2010). This could be due to better developed karst
as well. It is believed by some that the UCL which is the host rock of the perched
aquifer is more karstic than the LCL however one should also keep in mind that the
UCL has no cover. Large karstic features in the LCL such as the cave Ghar Dalam
show that this rock is also subject to carbonate dissolution (Stuart et al.., 2010).

Page | 24

Methods

Geologic control factors are important in the understanding of groundwater. The


understanding of the hydrogeology of Malta requires a huge amount of information
especially when considering its regional area. This study is therefore subdivided in
smaller sections which facilitate the ability to get an insight from several points of
view.
A one week field trip to Malta was carried out towards the end of June 2014.
Collected field work data includes geomorphologic site reconnaissance and
discontinuity scan line data collection. A number of sites were visited (Figure 11
adapted from the Geological Map of Malta, 1993). The site names and locations are
cross-referenced to the site reference numbers indicated on the map (Table 5).

5km

Figure 11 Sites visited indicated on the Geological Map of Malta (1993). Sites marked in red include only a site
reconnaissance exercise while magenta sites include also the collection of discontinuity data. Discontinuity data
for site 8 was acquired and not carried out by myself. No site visit was paid to site 8. Site numbers are crossreferenced to Table 5.

Page | 25

Site Ref.

Site Name

City/Village

Qammiegh

Mellieha

L-Imgiebah Bay

Mellieha

Fomm ir-Rih Bay

Bahrija

Wied l-Isperanza

Mosta

Wied il-Ghasel

Mosta

Gharghur

Gharghur

St. Georges Bay

St. Julians

Birkirkara Bypass

Birkirkara

Tal-Qroqq Area

Msida

10

Xghajra coastline

Xghajra

11

Munxar Area

Marsascala

Table 5 Table showing site names of sites included in study

Previous data that was acquired during the course of this study is re-interpreted
and used. This data set includes investigation borehole logs (Gianfranco et al., 2003),
well pumping tests determining aquifer transmissivity and potentiometric data
(BRGM, 1991b; BRGM, 1991c). It also includes field mapping of faults found in the
literature defining fault architecture style and predictive equations for the
development of fractures in fault zones (Faerseth, 2006; Bonson et al., 2007; Michie
et al., 2014; Missenard et al., 2014).
In this chapter all the methods used are explained.

3.1

Geomorphologic Site Reconnaissance

A number of sites to visit were identified on the basis to include a wide variety of
possible observations. The criteria include access to the main formations of Malta,
fault zones and not, inland and coastal areas. Each site is presented by using an
aerial photo or map on which the site photos are cross-referenced by numbers and
points of view for orientation. Geomorphologic features observed are highlighted

Page | 26

by annotations both on the aerial photo/map and on the site photos and are
discussed in the text. Some features such as mapped faults may be extracted from
the Geological Map of Malta (1993) if site observation is unclear.

3.2

Dip (angles) and dip directions of discontinuities

The aim is to identify possible trends and variations of discontinuity data. This data
was collected from four sites chosen at random and along horizontal orientations,
with another sites data acquired from external sources (Figure 11). A summary of
the data collected is presented (Table 6). Some of the rock faces studied are not
referenced to scan line positions and are therefore indicated as random (Table 6),
however for convenience this distinction is not made throughout this study. The dip
and dip direction of the discontinuities were measured by a geological compass of
the Brunton Geo Transit 5010 type.

Site
3
7

Fomm irRih Bay


St.
Georges
Bay

Msida

10

Xghajra

Birkirkara

(S)can Line
or
(R)andom
S
S
S
S
S
R
R
R
R
S
S
R
R
R

Ref.
A
B
A
B
C
A&E
B
C
D
A
B
A
B
E

No. of
readings,
n
15
7
95
73
55
18
15
2
11
19
22
29
13
4

Length
(m)

Normal
to face

7.5
6.6
33
20.6
14.7
42.1
23.3
6.3
36.2
18.2
11.3
-

N310
o
N325
o
N088
o
N340
o
N290
o
N138
o
N045
o
N005
o
N285
o
N240
o
N240
-

Formation

Personnel

L.G.L.
M.G.L.

CS

L.C.L.

CS & AM

L.G.L.

CS & AM

L.G.L.

CS

L.G.L.

AM

Table 6 Summary of discontinuity data collected. Site reference numbers are cross-referenced to Figure 11 and
Table 5. Initials CS refer to Christian Schembri and AM refer to Geotechnical Engineer Adrian Mifsud.

One should note that for Fomm ir-Rih Bay and Xghajra the scan lines orientations
are approximately the same therefore representing only one of the threedimensional directions per site. For St. Georges Bay and Msida two from the three
dimensions of space are represented. The scan lines orientations for Birkirkara site
are unknown. BS 5930:1999 suggests that discontinuities data is to be collected at
Page | 27

three orthogonal orientations. This is believed to represent the three dimensions of


space and thus limiting any data bias to a minimum. This is not achieved by the data
set presented here. The data may also be biased by not capturing the bedding
which from the regional geology is known to be generally sub-horizontal.
Joint Set

Site /
Scan
Line

3/All

21

70/359

80/255

3/A

14

89/005

3/B

7/All

223

7/A

J5

J6

(bedding)

(bedding)

79/255

86/356

66/259

11/057

95

86/356

58/237

12/060

7/B

73

80/353

84/192

68/258

9/314

7/C

55

68/158

70/217

89/031

20/045

74/117

9/All

46

90/337

90/157

9/A&E

18

84/340

61/113

9/B

15

90/336

90/156

9/C

9/D

11

10/All

41

85/325

85/149

5/056

10/A

19

87/144

13/090

10/B

22

83/324

83/152

5/045

8/All

46

25/174

8/A

29

22/180

8/B

13

35/165

8/E

J1

J2

J3

J4

J8

J7

(bedding)

Table 7 Summary of the main identifiable joint sets from pole concentrations for each site and scan line. Site
reference numbers are cross-referenced to Figure 11 and Table 5.

Discontinuity data is plotted on an equal angle lower hemispherical stereonet


projection separately for each scan line and for combined data for each site
(Appendix D). The software Dips v.5.1 was used for this purpose (Rocscience, 2004).
On the stereonet plots each discontinuity is represented by a pole and an overlaid
contour plot shows the pole concentrations from which the main joint sets are
identified (Table 7). The main joint sets are shown both as numbered poles and
Page | 28

planes on the stereonet plots (Appendix D). The variation across the sites of the
main joint sets is summarised (Table 8). These brackets represent the variability of
averages and therefore they do not portray all the variability. The joint set brackets
were widened to include all discontinuity data (Figure 12) so as to be able to
analyse and compare joint set characteristics. For this scope two other joint sets
being J9 and J10 are added. The latter joint sets never appeared as main join sets in
any of the sites.

Joint Set
J1
J2
J3
J4
J5 (bedding)
J6 (bedding)
J7
J8

from
70o
68o
58o
89o
5o
9o
61o
22o

Dip
to
90o
90o
80o
89o
20o
9o
74o
35o

Range
20o
22o
22o
15o
13o
13o

from
N324o
N144o
N217o
N031o
N045o
N314o
N113o
N165o

Dip Direction
to
Range
o
N005
41o
N192o
48o
o
N259
42o
N031o
N090o
45o
N314o
o
N117
4o
o
N180
15o

Table 8 Summary of the main joint sets variability of averages

Figure 12 Stereographic projection showing the entire joint set brackets considered for grouping of all the
discontinuity data overlaid over a scatter plot of all the pole data. The inner circle represents a dip angle of 35o;
poles within it represent discontinuities with dip angles less than 35o. J9 & J10 are introduced to include all of
the data.

Page | 29

3.3

Other discontinuities characteristics

Other discontinuity characteristics including persistence, termination, aperture,


infill, roughness, shape, wall strength and seepage were also collected for a good
number of discontinuities (Appendix E). This data set was collected by a visual
inspection aided by the use of a geologic hammer for wall strength.
In Chapter 4, data is analysed on the basis of aperture width classes (Table 9) and
persistence classes (Table 10). These classes are based on BS 5930:1999.
Aperture width class
1
2
3
4
5
6
7
8

Aperture size
Very tight
<0.1 mm
Tight
0.1-0.25 mm
Partly open
0.25-0.5 mm
Open
0.5-2.5 mm
Moderately wide
2.5-10 mm
Wide
1-2.5 cm
Very wide
2.5-10 cm
Extremely wide
0.1-1 m

Table 9 Aperture ranges for each aperture width class. These aperture width classes are used in the
presentation of Appendices F and G graphs. These are the basis for the analysis presented in Chapter 4.

Persistence
class

Persistence

Very low
(<1m)

Low
(1-3m)

Medium
(3-10m)

High
(10-20m)

Very high
(>20m)

Table 10 Persistence ranges for each persistence class. These persistence classes are used in the presentation
of Appendices F and G graphs. These are the basis for the analysis presented in Chapter 4.

3.3.1.

Relative hydraulic conductivity

An equation to describe the hydraulic conductivity as a function of average aperture


(e) and frequency () of joints for the ith joint set was developed in the literature
and is given below (Snow, 1970).

Page | 30

Ki =

ei3.i.g

12v
where subscript i refers to the ith joint set
K is the hydraulic conductivity
e is the average aperture of joints
is the frequency of joints
g is the acceleration due to gravity
v is the fluid viscosity
The frequency of joints () is calculated by dividing the number of readings (n)
within each joint set per site by the total scan line summed up lengths per site. The
average aperture (e) is calculated for the median value of each joint set population
by taking the average aperture of the range within which the median value lies
(Table 9, Appendices F & H). Through this method the measurement of aperture is
not accurate however it provides a good qualitative assessment at this stage.
A qualitative assessment of the relative K by comparing ei3.i of each joint set per
site is carried out (Appendix H). Therefore for the scope of this study the relative K
is understood as the term ei3.i.

3.4

Transmissivity

An extensive set of transmissivity data from a well pumping test campaign carried
out between January and September 1990 is used for re-interpretation (BRGM,
1991b). An analysis of transmissivity variation against depth of well boreholes ends
below the mean sea level, below the top of the LCL and proximity to the nearest
fault is carried out. The data used including data in graphical and summarised
tabular form is presented (Appendix I).
This study cannot confirm or otherwise the correctness of the transmissivity data
since the data available at hand is limited however it is still useful to provide a
background to report by BRGM (1991b).

Page | 31

BRGM (1991b) uses the interpretation methods of Theis, Hantush and Gringarten
with the most widely used method being the Theis method. The Hantush method
was used for those wells partially penetrating the aquifer and the Gringarten
method was used to model vertical fractures. Specific data about any of the test
configurations is not available. The Theis method is not usually used for unconfined
aquifers. Given the small drawdowns, which vary between 0 to 15 metres and are
only up to 10% of aquifer assumed thickness, the Theis method is deemed fit for
purpose. Changes in water level are deemed to be due to pressure change as the
Theis method requires and not by gravity. In fact BRGM (1991b) confirm that the
field data fits well with the Theis master curve.
The geologic reasons behind the variability of transmissivity data may be various.
These may include variations in fracture population, karst processes, permeabilities
of lithologies and facies and possible seals due to faults. With the data available at
hand it is not possible to try and test the effect of each of these factors. Detailed
borehole logs including both geological and geotechnical data would have been a
good starting point for this scope. However hints of causes of variability are
searched for by the analysis carried out.

3.5

Potentiometry

BRGM (1991c) have drawn potentiometric contoured maps for the years 1944, 1969
and 1990. The most reliable set of data is that collected between 1988 and 1991.
Defects in the previous data sets include measurements not taken frequently,
surveying errors, wrong or missing calibration of equipment and unidentified
geologic occurrences such as sparse presence of clay lenses which produced local
potentiometric highs and were not excluded.
The potentiometric map of 1990 is thus presented together with a brief discussion.

Page | 32

3.6

Control of fault parameters on hydraulic properties

The aim is to develop an understanding of the control of faults on the hydraulic


properties of Maltas geologic environment. The complexity involved in devising a
methodology with this aim is highlighted. As this study includes only a very limited
data set acquired through first-hand field work, which is not always related
specifically to this part of the subject, existing data and relationships are used in
developing our understanding. It should be understood that such data set although
useful is not exhaustive and has various limitations.
As the ENE-WSW faults are the main fault outcrops of Malta the Victoria fault is
chosen as a case study. As the major rifting occurred at the end of deposition, it can
be assumed that the whole stratigraphic sequence was affected. The Victoria fault
zone architecture is best preserved at the coasts.
It is widely known that hydraulic properties of rocks are governed both by their
matrix properties and the fracture distributions within them. In this section a
plausible desk study type methodology to characterise the Victoria fault zone and
thus fracture distributions is presented. Fault parameters are presented in each
sub-section followed by an explanation highlighting their importance for our scope,
the methods, difficulties and assumptions taken.

3.6.1.

Displacements along fault lengths

If displacements at particular points of the fault are known predictive equations can
be used to predict widths of fault zones, total damage zones, fault cores and
fracture population (Michie et al., 2014). Displacement-length relationships provide
a tool to interpolate for displacements along faults where displacement data is not
available.
Fault growth models describing segmented isolated faults have been proposed
following the relation D = cL2 (Watterson, 1986; Cartwright et al., 1995). Knowing

Page | 33

the displacement (D) at three points along the length of the fault (L) would allow an
approximation of this form. A major uncertainty lies in the verification of the
constant c which relates to a number of fault and material properties (Watterson,
1986).
Assuming that the Victoria fault is a segmented isolated fault, the indicated coordinates of D and L are used to obtain a formula in the form above (Figure 13;
Table 11; Appendix J).

N
3

2
3km

Figure 13 Points of which the details are presented in Table 11. The red line indicates the direction along
which the length of the fault is considered (adapted from The Geological Map of Malta, 1993)

Point Ref.
1
2
3

D (m)
150.2
205.4
(assumed as
max D)
60

L (m)
-1700
0
10400

Reference
Figure 11 of Dart et al. (1993)
Geologic Section A-A in the Geological
Map of Malta (1993)
Michie et al. (2014)

Table 11 D and L co-ordinates used to determine a relationship in the form D=cL2

The resulting approximation is D = 1.3x10-6 L2 with c about two orders smaller than
the values suggested by Watterson (1986). With this equation the full length of the
fault approximately equals 26 kilometres. If c being small is incorrect it might show
that the assumption of the Victoria fault being a segmented isolated fault is wrong.
The assessment of inter-linked overlapping faults is more complex since the lengths
of each segment would need to be measured and interpreted (Figure 14 adapted
Page | 34

from Cartwright et al., 1995). Given the current lack of displacement-length data
along Victoria fault zone this formula cannot be used to predict displacements at
points of this fault.

Figure 14 Graph showing growth path of a segmented isolated fault and the complexity to determine this
relationship due to field data scatter. For inter-linked overlapping faults the difference between the green point
and the blue point depends on what fault length is considered. (adapted from Cartwright et al., 1995)

3.6.2.

Length and termination points of faults

The importance of faults length data in determining displacement-length


relationships has been highlighted.
Several difficulties to determine end points of faults and thus their length may be
encountered. Most of the mapped faults of Malta have at least one part reaching
either west or east coast with the Victoria fault reaching both. Offshore faulting to
the east may lie below seismic resolution meaning that fault with throws less than
10 metres are difficult to map (Dart et al., 1993). A lack of good fault inland
outcrops due to back scarp erosion may have erased some fault termination points.
This same reason makes it difficult to use aerial photography for this scope. On the
other hand, some quarries located along Victoria fault length may provide useful
information.

Page | 35

The Geological Map of Malta (1993) suggests en-echelon left stepping fault
structures at the east onshore stretch of the Victoria fault. Detailed field mapping of
these may provide some useful information on length and displacements of faults.
However given the complexity of this problem and the current lack of data the
relationship in the previous section is not developed further. If we consider using
the fault data at point 2 (Figure 13) from the Geological Map of Malta (1993) and
assuming that the geometry proposed by Michie et al. (2014) can be applied at the
scale of the Victoria fault we expect a fault splay zone width of about 200 metres
(Appendix J). This coincides well with the UCL syncline due to fault drag observed at
Fomm ir-Rih Bay. With the current displacement-length relationship the fault zone
width would be expected to gradual narrowing towards the east coast.

3.6.3.

Fault architecture at a cross-section

The predictive equations proposed by Michie et al. (2014) have a potential. If they
are combined with knowledge of stratigraphy and displacements they may provide
useful predictions of the fracture populations to expect at a cross-section of the
ground. However these equations have their limitations too. Limited detailed
information is available for displacements larger than 25 metres and therefore the
application at these ranges needs further testing.
In proposing a conceptual ground model the average fault dip angles proposed by
Michie et al. (2014) are assumed to be constant for all displacements.
The following is an example of how from heave data, damage zones widths may be
calculated. Cumulative heave plots show that the maximum heave on a single fault
is approximately 10 metres at a stretch south of the Victoria fault (Figure 15
adapted from Putz-Perrier & Sanderson, 2010). An approximate displacement of 36
metres can be calculated from simple geometry if an average fault dip angle of 74o
is considered (Michie et al., 2014). Using the predictive equations proposed by
Michie et al. (2014) gives us the maximum developed widths of fault splay zones,
Page | 36

total damage zones and fault cores (Table 12; Appendix J). However these depend
on having an available thickness of the relevant geologic strata above the point
where this displacement has occurred.

Figure 15 Cumulative heave plot for a cliff scale section named Malta D with location shown on map. Red lines
on plot indicate single faults having major heave spaced at approximately 600 metres. (adapted from PutzPerrier & Sanderson, 2010)

Formation
GL
LCL

Fault Splay Zone


Width (m)
22.2
2.1 15.3

Total Damage Zone


Width (m)
71.2
14.3

Average Fault Core


Width (m)
0.6

Table 12 Expected maximum widths for a 36m displacement using predictive equations from Michie et al. (2014)

3.6.4.

Strata thickness and properties

Variable behaviour is expected from different strata due to their different


mechanical properties. Since strata behave differently, their thickness is important
to know.
The readily available LCL structural contour map, isopach maps and topographic
levels are used (Pedley, 1975; the Geological Map of Malta, 1993). A series of crosssections are drawn across the Victoria fault, which are used in the construction of
the regional conceptual model.

Page | 37

Observations and Analysis

The observations and the analysis from the methods and the data highlighted in
Chapter 3 are detailed in this section.

4.1

Geomorphologic Site Reconnaissance


4.1.1.

Qammiegh

Figure 16 Aerial photo of Qammiegh Site (adapted from Google Earth, 2014). The red polylines encircle slope
instability areas which mainly involve rock toppling. The blue polyline encircle a wetland and saltmarsh. Black
lines show the locations of faults and where they are dashed it means that they are inferred (The Geological Map
of Malta, 1993). Numbers are cross-referenced with photo numbers of this section and the arrows show the
orientation of view.

From the aerial photo of Qammiegh (Figure 16 adapted from Google Earth, 2014)
one can notice that the width of the rock toppling deposit area varies along the cliff
length of UCL. At the area north-west of Qammiegh Fault, one can notice that at the
part closest to the fault, the deposit zone does not extent to the coast. North of the
Qammiegh fault, the general dip of the UCL and BC strata is towards the north-east
(The Geological Map of Malta, 1993). To the west the deposit area widens gradually
Page | 38

as the slope direction is approaching the dip direction of the strata. Further to the
north-east the deposit area narrows again as the topography lowers and thus relief
is smaller.
At the rock toppling area south of the Qammiegh Fault, rock toppling can only be
observed at that stretch of coast where the BC outcrops above sea level. Whether
this phenomena of rock toppling continues below sea level or whether its
manifestations occur at a slower rate is unknown. It might be plausible to think that
any of the latter statements hold since rock toppling is associated with the different
behaviour of the BC than the LCL under wetting and drying cycles. In cases where
the BC is constantly saturated, the situation might be different from what is
observed at slopes with outcropping BC above sea level.

Figure 17 Photo 1. Arrows indicate extent of main karst features and dashed lines show enlarged joints
observed from distance.

Extensively developed karst is observed in the UCL (Figure 17). Karst features are
observed to be dominant either in the vertical or the horizontal direction closer to
surface and their existence seems to be related to wider joints. Some karst features
which do not have either of the two axes as dominant are also observed both close
and away from surface. Karst at this level together with the different mechanical
and hydrogeologic properties of UCL and BC can be a trigger of cliff retreat. An

Page | 39

interpretation might be that it may have developed extensively at this level due to
different eustatic levels or at some early stages during the uplift of Malta.

Figure 18 Photo 2 showing extensive cave development at Fault Zone. Dashed line indicates inferred fault (The
Geological Map of Malta, 1993).

Figure 19 Photo 3 showing UCL hanging wall

An interesting occurrence is that of extensive cave karst development at a mapped


fault close to sea level (Figure 18). Karst development predominantly at the bedding
discontinuities is also observed (Figure 19).

Page | 40

Varying slope angles along a slope section comprising the main four formations of
the Maltese Islands are observed (Figure 20). Shallower slope angles occur for finergrained strata with the steepest angles occurring for larger-grain dominated strata.
Michei et al. (2014) observes that fault dip angles within finer-grain dominated
geologic strata are shallower than for coarse-grain dominated strata. This occurs at
Qammiegh even for slope angles.

Figure 20 Photo 4 showing varying slope angles along a slope section. Red lines indicate general slope angles
of every formation.

4.1.2.

L-Imgiebah Bay

Zones along the UCL BC contact are observed to be vegetated (Figure 21 adapted
from Google Earth, 2014). These can be indications of presence of springs at this
contact. Variability in the vegetation density is noted, with the west part being
more densely vegetated than the east part. A larger relief at the west part is noted
due to an approximately double thick UCL stratum at the west part when compared
to the east as observed from the structural contours of The Geological Map of
Malta (1993). The west part therefore also has a higher water storage capacity
which translates into more flow on this side of the valley.

Page | 41

Rock toppling phenomena are more evident towards the coast which can be due to
greater relief at these locations. Lack of vegetation is noted at these areas when
compared to the more inland areas.

Figure 21 Aerial photo of L-Imgiebah Bay (adapted from Google Earth, 2014). The red polylines encircle slope
instability area. The green polylines encircle denser vegetated areas. Numbers are cross-referenced with photo
numbers of this section and the arrows show the orientation of view.

The angles observed here are steeper (Figure 22) for both formations when
compared to those at Qammiegh (Figure 20), however before hypothesising of what
can be the cause for this difference one should also keep in mind that there is an
order of magnitude of difference in the scales of observation. Steeper angles for the
UGL occur with its face being sub-vertical. At least three families of joints are
observed (Figure 22). These include approximately oriented ENE-WSW joints, NWSE joints and the sub-horizontal bedding. UGL blocks between the former two joint
families are noted to have made their way down. A plausible interpretation is that
water seeps down through desiccated clay surfaces (Figure 23) and follows paths
along UGL joints widening them and stressing the UGL until UGL blocks are
detached completely.

Page | 42

Figure 22 Photo 1 showing cliff edge with the BC overlying the UGL. Dashed lines indicate joints into the
plane of the paper (approx. ENE-WSW), dotted polylines indicate joints parallel to the plane of the paper
(approx. NW-SE).

Figure 23 Photo 2 showing desiccated clay surface


(scale shown by field notebook scale line)

Figure 24 Photo 3 showing oxidised UGL joint as


interpreted by Missenard et al. (2014)

Page | 43

An oxidised UGL joint located at the base of this cliff section is observed ( Figure 24).
This can be interpreted as a palaeo-fluid flow path in the sense as analysed by
Missenard et al. (2014). In such a case this can provide supporting evidence in a
temporal framework to the hypothesis here proposed, in that seeping water
through discontinuities has been responsible for cliff retreat of this section.
Walking over UGL sub-horizontal surfaces it is observed that joints are tighter and
have larger spacing away from the cliff face. Seeping water may be less at more
inland locations and if any, the side stresses imposed on rock blocks would be
supported by adjacent blocks. Rock blocks at the edge of the cliff face do not have
adjacent rock blocks to provide this support and thus the discontinuities here are
susceptible to widen at a larger rate.
Similar observations of joint families that were observed in the UGL (Figure 22) are
also observed in the UCL (Figure 25). From what can be observed from a distance it
seems that the joints have a wider aperture and are more karstified in the UCL .

Figure 25 Photo 4 showing UCL cliff edge exhibiting at least two dominant discontinuity sets. For explanation
of annotations used refer to Figure 22.

Page | 44

4.1.3.

Fomm ir-Rih Bay

Figure 26 Aerial photo of Fomm ir-Rih Bay (adapted from Google Earth, 2014). Numbers are cross-referenced
with photo numbers of this section and the arrows show the orientation of view.

On the way to Fomm ir-Rih Bay, similar karst features to Qammiegh (Figure 17) are
observed at the UCL in the nearby ridges. The aerial photo of Fomm ir-Rih Bay is
presented (Figure 26 adapted from Google Earth, 2014).

Figure 27 Photo 1 showing Victoria fault at west coast outcrop. Annotations used are explained by notes on
photo.

Page | 45

A photo of the Victoria fault is presented (Figure 27). The occurrence of smeared
clay is not clearly observable at this location due to vegetal cover. In the literature it
is reported that this part of the fault has a good seal by clay smear due to large
throws (Sapiano et al., 2006; Missenard et al., 2014). An antithetic fault to the main
fault is observed (Figure 27; Figure 28). Two zones of sub-vertical joints are
observed close to this fault which may form part of the Fault Splay Zone which
widens in fine-grained materials (Michie et al., 2014). An LGL outcrop is observed to
have rounded joint profiles (Figure 29). Black staining from some of the joints is
interpreted as an indication of recent fluid flow. The MGL member is more marly in
nature. An MGL outcrop exhibits sub-horizontal dessication patterns which may be
due to sedimentation processes (Figure 30). Complex networks of secondary cracks
between main sub-vertical discontinuities probably as a result of the shear zone are
observed.
An indication of the strength by the use of a geologic hammer shows that the MGL
is weaker than the LGL (Appendix E). A more detailed assessment requires at least
the use of a Schmidt hammer. The Lower Main Phosphorite Conglomerate (LMPC),
a marker between the LGL and MGL, is noted (Figure 31; Pedley & Bennett, 1985). It
overlies a hardground, which is similar to a desiccated crust. Probably the erosion
resistance of the latter is higher than that of the LMPC given that the LMPC was not
noted close to the cliff edges but mostly in more protected zones. Hardgrounds are
therefore expected to have lower permeability.

Page | 46

10m

Figure 28 Photo 2 showing detail of antithetic fault shown in


Figure 27.

Figure 29 Photo 3 showing LGL outcrop at


location of Scan line A (scale shown by geologic
hammer)

Figure 30 Photo 4 shows MGL at location of scan line B.


Dashed lines show examples of sub-horizontal desiccation
joints (scale shown by field notebook)

Figure 31 Photo 5 showing the Lower Main


Phosphorite Conglomerate (scale shown by field
notebook scale line)

Page | 47

4.1.4.

Wied il-Ghasel

Figure 32 Karst caves in the Attard member of the LCL at Wied il-Ghasel where it cross-cuts the Victoria fault,
the rock face strike is approximately NNE

Extensive karst caves development can also be observed in the LCL (Figure 32).

4.1.5.

Gharghur

An aerial photo of the site visited at Gharghur is presented (Figure 33 adapted from
Google Earth, 2014). A strip of land indicated by a green line has denser vegetation.
When compared with the Geological Map of Malta (1993) this line seems to follow
an inferred contact between the Xlendi (upper) and Attard (lower) members of the
LCL. This fits well with the occurrence of karstic caves such as Ghar Hassan and Ghar
Dalam at the south of Malta which are known to be developed between il-Mara
member (over Xlendi member) and Attard member (BRGM, 1991b). The Attard
member is documented to have a north-south fine-grained facia close to this
location (the Geological Map of Malta, 1993). Due to larger flows in the coarsegrained strata it is expected to have higher erosion development in the Xlendi
member.

Page | 48

Figure 33 Aerial photo of site visited at Gharghur (adapted from Google Earth, 2014). The green line shows a
stretch of denser vegetation. Numbers are cross-referenced with photo numbers of this section and the arrows
show the orientation of view.

Notches at the Xlendi member of the UCL are noticed (Figure 34). This coincides
with a wide aperture joint at which flow can be witnessed to be higher by the
observation of black staining. Tighter joints are observed above this level at the
same cliff face but are not black stained.
A detail of calcite deposition is presented (Figure 35). The rock face takes shape
similar to that of flowing water. This may be an indication that either flow paths
have been long enough for water to become saturated with calcium carbonate in
relation to the host rock, that water flow is slowing down and losing energy and
thus depositing calcite or more likely a combination of both. Black staining is also
observed here which may indicate recent flow pathways. A slight different karst
feature is observed at a karstified cave pillar which seems to have much less calcite
deposition (Figure 36).

Page | 49

Figure 34 Photo 1 showing formed notches at Xlendi member of the LCL. Dashed polyline indicates a wider
aperture joint.

0.5m
Figure 35 Photo 2 showing detail of calcite
deposition

1m
Figure 36 Photo 3 showing karstified cave pillar

Page | 50

4.1.6.

St. Georges Bay

Figure 37 Infilled joint having


wide aperture
(scale shown by tape measure)

Figure 38 Fault Breccia


(scale shown by car key)

Figure 39 Wide open aperture at


fault zone
(scale shown by geologic hammer)

Some interesting observations are highlighted from joints and fault zones studied at
St. Georges Bay. The formation here is the LCL and more specifically the Xlendi
member.
A wide joint, at places reaching apertures in the region of two centimetres infilled
with terrarossa traces and zones of calcited walls is shown (Figure 37). This joint
shows at least past fluid flow however at such a state present fluid flow is more
difficult to occur. In such a state fluid flow, if any, would be encouraged to find
easier and wider routes. Such a situation probably results in deviation of fluid flow
paths. Two scenarios are possible in this case either infilling of other joints or
karstification of others. It is believed that this is a function of joint characteristics
such as orientation, aperture, persistence and frequency.
A zone of fault breccias is observed (Figure 38). The cataclasite is angular with some
rounded corners which could have formed during deformation given that the rock is
not very hard. The soil matrix is lighter in colour than that observed in Figure 37 and
also has a good proportion of coarse-grained particles when compared with the
latter. These occurrences can be an indication that at least some of the soil matrix
formed during deformation. The properties of this cataclasite are expected to affect

Page | 51

largely the permeability across and along this joint. Some important properties to
consider are particle size distribution, voids ratio and particle orientation.
Different fault zone architecture with a rock block occupying it is observed (Figure
39). This rock block became dislodged and broken with no evidence of fault breccia.
It is plausible to believe that the fault displacement in the latter case is less than
that for the fault seen in Figure 38 where fault breccias did develop to
accommodate deformation. Unfortunately these interpretations cannot be backed
up by site observations since the height of the outcrop is very short due to past
excavation. In this zone fluid flow is expected to be encouraged through due to
wider aperture.

4.1.7.

Msida

The main observations at the Msida site visited are highlighted. A schematic plan
layout of the site is presented (Figure 40). Only minor karst features are observed in
the LGL member (Figure 41). In general the karst features observed in any of the GL
members were not extensive as observed either in the UCL or LCL members. The
major karst feature observed at this site involves a vertical karstified discontinuity
of about 1.5 metres in height (indicated in Figure 44). The discontinuities observed
at the faces of this site are generally grouped in zones which are spaced at between
four and eight metres (Figure 42, Figure 43 and Figure 44). The apertures are
generally tight.

Page | 52

Figure 40 Schematic plan layout of Excavation site


visited at Msida with an indication of the excavation
faces references

Figure 41 Minor Karst feature noted at wall A


(scale shown by pencil and Barton comb)

Figure 42 Excavation face A with dotted lines indicating discontinuity zones (scale is shown by excavator)

Page | 53

Figure 43 Excavation face B with dotted lines indicating discontinuity zones (scale is shown by excavator)

Figure 44 Excavation face D on the left side, face B on the right side and face C in between. Discontinuity
zones are indicated by dotted lines (scale is shown by excavator)

4.1.8.

Xghajra

NE striking discontinuities can be widely observed along this part of the coast
(Figure 45 adapted from Google Earth, 2014; Figure 46). The discontinuity at the LGL
wall has a shallower dip angle than that observed in the LCL and thus observations
by Michie et al. (2014) do also apply here. Stretches along the length of the

Page | 54

discontinuities at the LCL outcrop can be observed to be widened by fluid flow and
the surface of the same member is everywhere lightly karstified. This occurrence
was also observed at other UCL outcrops such as at l-Imgiebah. It is plausible to
believe that karst develops mostly at locations of a larger fluid flow.

Figure 45 Aerial photo of site visited at Xghajra (adapted from Google Earth, 2014). Dashed lines show
examples of NE discontinuities which can be observed even from this aerial photo.

An interesting observation at Xghajra is the extremely wide aperture at the contact


between the LGL member and the Il-Mara member of the LCL formation which is in
the region of 200 millimetres (Figure 46; Figure 47). This gives us a good reason to
agree with the interpretation by BRGM (1991b) if this contact is located below the
water table it acts as a fluid conduit.
Remains of the Scutella echinoid bed on top of the LCL forms the transitional bed
between Il-Mara member of the LCL and the LGL (Figure 47; Pedley, 1975).

Page | 55

LGL

LCL

Scutella echinoids
Figure 46 Photo 1 showing NE trending
discontinuities on the LGL wall and on the LCL
ground (scale shown by field notebook. Dashed lines
show discontinuity dip angles. The polyline shows a
karstified stretch on the LCL ground.

4.1.9.

Figure 47 Photo 2 showing contact between LGL


and LCL and Scutella echinoids marker (scale shown
by field notebook)

Munxar

The outcrop at the site of Munxar is the MGL (Figure 48). A morphologic difference
between this member and the other two members of the GL can be observed from
the style of surface sub-horizontal discontinuities similar to those observed at
Fomm ir-Rih Bay but better seen here (Figure 49). They seem to be desiccated
discontinuities related with sedimentation processes.
The marly MGL cliff face is very friable (Figure 50).

Page | 56

Figure 48 Aerial photo of Munxar Site (adapted from Google Earth, 2014)

Figure 49 Photo 2. Dashed lines show examples of tight discontinuities which may be interpreted as
sedimentation desiccation discontinuities. (scale shown by field notebook scale line)

Page | 57

Figure 50 Photo 1 showing a low MGL cliff face

4.2

Inferring Contacts from Boreholes


4.2.1.

UCL/BC

300m

Figure 51 Investigation boreholes location encircled in red (source: Google Earth, 2014)

Page | 58

It was difficult to closely inspect the UCL/BC geologic contact at the sites visited due
to several toppled rock blocks and debris. For this reason a set of investigation
borehole logs are studied to observe this contact (Appendix C). The set of boreholes
included a set of 12 from a location at the old citadel Mdina (Figure 51; Appendix C).
The description included a sandy material at the UCL/BC contact. This can be either
transported material from the surface through the fracture network or the
greensand material which is reported to be a very thin layer between the UCL and
BC at certain areas.

4.3

Dip (angles) and dip directions of discontinuities

The stereonet plots on which this section is based are attached in Appendix D.
4.3.1.

Fomm ir-Rih Bay

The joint sets identified at this site, J1 and J3, closely resemble the two main fault
families of Malta, the ENE-WSW and the NW-SE trending faults respectively.
Interestingly J1 was identified at this site even though the outcrops almost lie at the
footwall of the Victoria fault (ENE-WSW trending). This unbias of the data is
probably the result of the approximate angle between the orientation of the rock
face of the scan lines and the strike of the joint sets being at least 45o. The average
strike of J1 at this site is E-W which closely resembles the orientation of the Victoria
fault at this location. A trend of the bedding is not identified.
At scan line B only few readings were taken due to the short length of the outcrop
at this point and due to the omission of the secondary fractures in between the
main discontinuities (Figure 52). Unarguably the secondary fractures, that form part
of the fault damage zone, increase the permeability of this zone. However they
were omitted from the data collection with the aim to try and identify main
discontinuities trend similarities between the different sites. The readings thus
taken at scan line B are only seven however when they are combined with the
readings for scan line A, 4 of the readings reinforce what is observed in scan line A.
Page | 59

2m

Figure 52 MGL outcrop at Fomm ir-Rih Bay scan line B. Dashed lines show main joints.

4.3.2.

St. Georges Bay

A wide variability of discontinuity data is noted at this site. This can be due and the
close proximity of the site to a mapped fault (the Geological Map of Malta, 1993).
Due to previous excavation works, it is not always easy to identify between main
structural discontinuities and secondary discontinuities. This problem was tackled
by taking a large sample so as to be able to identify the main trends. After
combining all of the data for this site the three main discontinuities are identified.
The most predominant joint set is J5 (bedding) with the dip direction in close
agreement with the mapped direction (the Geological Map of Malta, 1993). The
second most predominant joint set is J1 with an average strike being E-W trending.
This resembles more the strike of the Victoria fault at the west half of Malta rather
than the closest exposure of the Victoria fault which trends approximately ENEWSW. The mapped Victoria fault is shown to have a general kink west of Mosta
from a close to an E-W orientation to an ENE-WSW direction (the Geological Map of
Malta, 1993). The third most predominant is J3 which strikes approximately NNWSSE at this site. Therefore the two main fault families are also well represented at
this site.

Page | 60

If the scan lines are to be analysed separately it is less evident which joint sets are
the most predominant, however one can note that the highlighted joint sets J2 and
J4 could be antithetic to J1 and J3 respectively. This occurrence closely reflects the
structural geology of Malta with its horst and graben structure. Two other joint sets
J6 and J7 are also observed. J6 probably refers to bedding which dips approximately
NW while J5 dips approximately NE. Being close to a fault zone this variability is
expected.

4.3.3.

Msida

The main joint sets identified at Msida are J1 and J2. The average strike of these
sets is closer to the ENE-WSW orientation which closely reflects the trend of a
mapped fault at approximately 1.5 km south of this site (the Geological Map of
Malta, 1993). The discontinuity sample was limited due to the very wide spacing of
discontinuity zones and thus no more clearly dominant discontinuity sets could be
identified. However there seems to be steeply dipping sparse discontinuities
trending approximately NNE-SSW followed by NW-SE trending joints. A trend of the
bedding is not identified.

4.3.4.

Xghajra

The main discontinuity sets identified at this site are J5 (bedding) and J1 and J2
which trend approximately NE-SW. The latter two joint sets have approximately the
same average strike but have opposite dip directions with dips that are sub-vertical.
They closely resemble a mapped faults in the vicinity (the Geological Map of Malta,
1993). No dominant discontinuity set sub-parallel to the NW-SE trending mapped
fault (the Geological Map of Malta, 1993) can be identified which probably is due to
the bias imposed by the scan line orientation being sub-parallel to this fault.

Page | 61

A wide dispersal distribution of the poles is observed in scan line A which starts off
very close to a cave-like structure (Figure 53) and undoubtedly has an effect on this
data set. This effect is not noticed at scan line B which was carried out
approximately 75 metres away from scan line A and this structure.

1m
Figure 53 Cave-like structure adjacent to the position of the start of scan line A

4.3.5.

Birkirkara

From all five sites for which discontinuity data is available the data for this site is the
most unclear from a regional tectonic point of view. The main joint set J8 identified
at this site was never identified in any of the other sites. The average dip is shallow
at 25o especially when considering that this joint set is not predominantly made up
of bedding discontinuities. In the other sites it was in general observed that bedding
discontinuities had a very shallow dip angle while the steeper discontinuities
included other joints. From all the sites at which discontinuity data is available, this
site is the furthest away from highly stressed fault zones (Putz-Perrier, 2008; PutzPerrier & Sanderson, 2010) and mapped faults (the Geological Map of Malta, 1993).
By looking at a photo of the site (Figure 54) a plausible interpretation might be that
these discontinuities may well be related with the isopach thickening of the LGL

Page | 62

which is reported to be by previous up-arching of this member (Pedley, 1975;


Pedley et al., 1976).

West

East

Figure 54 Photo of Birkirkara site showing undulating discontinuity (scale shown by mobile crane)

4.4

Other discontinuities characteristics


4.4.1.

Aperture

Tables showing the population of each joint set per site within each aperture class
are presented (Appendix F). Plots of aperture width class against both dip directions
and dip angles are plotted separately for all the sites, grouped as joint sets and for
the GL formation (Appendix F). The aim is to identify possible trends of wider
apertures. The results are not very clear however some kind of relation is hinted to.
A slight concentration of wider apertures for the E-W trending joint sets are noted
at Fomm ir-Rih Bay, St Georges Bay and Msida and for the N-S trending joint sets at
St. Georges Bay and at Msida (indicated by circles on graphs in Appendix F). Some

Page | 63

apertures seem to increase with increasing dip angle at St. Georges Bay and Msida,
however at St. Georges Bay there are some shallower dips with large apertures too.
Since bedding data is limited, no real comparison of bedding aperture may be made.
At Xghajra and Birkirkara there is no clear trend of wider discontinuities of a
particular joint set. It can be noted that at Birkirkara the apertures are generally on
the wide end of the spectrum.
By plotting the same data and grouping it according to joint sets (Appendix F), it is
noted that the major concentration of wider apertures lies at those joint sets
trending approximately E-W namely J1, J2 and J4. The bedding joint sets (J5, J6, J8
and J10) do not clearly show a trend for aperture. Apertures of the approximately
NNE-SSW trending joint sets J6, J9 and J7 tend to be on the smaller side of the
spectrum with only one joint exceeding aperture class 5. Combining the aperture
data together with the orientations of the latter joint sets may show that the
tectonic origin of the two main fault families of Malta has the least effect on these
joint sets.
No obvious trend could be identified for aperture width in the GL formation mostly
due to the variability across sites. Discontinuities with dip angles ranging from 25o
to 50o are dominated by wider apertures from Birkirkara. At the latter site apertures
are generally on the wider third of the spectrum.

4.4.2.

Persistence

Given that the majority of the joints have steep dip angles it is difficult to obtain a
qualitative assessment of persistence from rock outcrop walls with a limited height
such as those studied at Fomm ir-Rih Bay, St. Georges Bay and Xghajra. In such
cases judgement was carried out on what is possible to observe with the
consequence that persistence may be underestimated for some joints. It is
therefore no surprise that most persistence data lies within classes 1 to 3.

Page | 64

Scatter plots for persistence class against dip direction and dip angle, for aperture
width class against persistence class and joint set populations within each
persistence class for each site are presented (Appendix G). The following
observations are made.
At Fomm ir-Rih Bay the higher persistence classes occur for joint sets J1 & J3 having
an ENE-WSW and NW-SE approximate strikes respectively. Some joints with the
higher persistence lie within the joint set J4 but are on the border line of J1. This is
only a matter of grouping and therefore the latter joints may be considered to be
part of J1 too. The apertures of the highest persistent joints tend to be on the
widest end of the spectrum too.
At Msida the higher persistence classes occur for joint sets J1 & J2 which have an
ENE-WSW strike. They are the most occurring joint sets too and reflect one of the
main fault families of Malta. Some of the higher persistence joints lie within joint
set J7 but are a border of joint set J2 and thus may be considered as part of J2 in
this case. The apertures of the most persistent joints vary from very tight to very
wide.
At Birkirkara only three data points have persistence class greater than 3 (circled on
graphs in Appendix G). These joints have similar orientation and dip angles with an
approximate strike of an E-W direction. They lie within two different joint sets J2 &
J8 but at their common border line. These joints have the aperture width class 7
which is almost the largest aperture in this site.
Given the large sample size of St. Georges Bay relatively very few joints have
persistence greater than 3. They are mostly part of the main bedding joint set J5
however they do not have the wider apertures of this site. If we expect to have
higher persistent joints to have the wider apertures the latter occurrence may
highlight clearly the difficulty of this outcrop to gather correct persistence
information.
No clear trend can be identified at Xghajra.

Page | 65

4.4.3.

Relative hydraulic conductivity (K)

At Fomm ir-Rih Bay the highest relative K occurs for joint set J4, which has an
average strike approximately ESE-WNW at this site. The second highest relative K
occurs for J3 which strikes approximately NW-SE but is considerably less than J4.
This shows that the relative Ks at this site are tectonically controlled and are highest
for J4 since it has the closest similarity to the nearby Victoria fault. However it
should be reminded that this data set was simplified by not collecting what are
termed as secondary fractures at scan line B (refer to section 4.3.1).
At St. Georges Bay the joint sets J1, J2, J3, J4 and J5 all have similar relative K in the
range from 1.3 to 1.9 mm3/m. These joint sets include the strikes of the two main
fault families of Malta and the bedding.
At Msida all relative Ks for all joint sets are relatively very small due to low joint
frequencies and low average apertures however J1 has the highest relative K. This
again indicates a plausible tectonic control even though the site lies about 1.5
kilometres away from the nearest mapped fault.
At Xghajra the bedding joint set J10 was measured only once however it had a very
high average aperture and thus has the highest relative K. One should remember
the approximately 200mm open GL-LCL contact. J10 may have well been affected by
this contact. The joint sets J7 and J3 follow but have a considerably lower relative K
than J10. They strike approximately NW-SE and again indicate a plausible tectonic
origin.

4.5

Transmissivity

Even though there is a wide scatter of data, transmissivity seems to decrease with
increasing depth below the top of the LCL (Figure 55). Reasons may include the
closure of cracks at depth due to higher vertical stresses, the infill of discontinuities
Page | 66

with transported materials and less occurrence of karst at deeper levels. The
dispersal of data may be explained by knowing that the LCL top is well above mean
sea level at the west while not so at the east coast (refer to structural map of LCL in
Appendix J which is sourced from Pedley, 1975). From west to east members and
lithofacies are probably encountered by the different well tests so one cannot really
generalize for the LCL. For these reasons BRGM (1991b) re-grouped this data by
regions and made some interesting observations. The contact GL-LCL is interpreted
as a conduit of water when located below water and the range between Il-Mara and
Attard members of the LCL is noted to have better productivity (BRGM, 1991b).
Their observations fit in well with the geomorphologic features observed at Xghajra
(Figure 47) and Gharghur (Figure 33, Figure 34, Figure 35 & Figure 36) respectively.

Figure 55 Variability of transmissivity with borehole depth below top of LCL (line shown is the trend line)

If we consider transmissivity variability with respect to the mean sea level (Figure 56)
one can note that the dispersal of data is less pronounced. Very simply this tells us
that the borehole depths were in general aimed at close to the sea level, however it
may give another piece of information. Transmissivity decreases at a faster rate
with depth below the mean sea level than it does with the depth below the top of
the LCL. This may be an indication of preferential karst features forming close to the
mean sea level given that in mean sea aquifers one expects to have developed karst

Page | 67

near the saline-freshwater contact (Mylroie & Mylroie, 2007), however the basis of
this hypothesis would be very weak if based only on this piece of data.

Figure 56 Variability of transmissivity with borehole depth with respect to the mean sea level (line shown is the
trend line)

Transmissivity is also plotted in relation to proximity to the nearest mapped fault in


the Geological Map of Malta (1993) considered at surface (Figure 57) and at depth
(Figure 58). A good number of data points that have an exceptionally wide scatter
are identified and ignored from the fitting of a possible trend line. It may be that
faults have less of an effect on transmissivity at distances further away than
approximately 600 metres (data points above the red arrow in Figure 57 & Figure
58). Some comparatively low transmissivity values occur relatively close to faults
too (shown circled in Figure 57 & Figure 58). If one ignores these values, there
seems to be a tendency of decreasing transmissivity with distance away from a fault.

Page | 68

Figure 57 Variability of transmissivity in relation to distance away from fault considered at surface. Black line
shows a possible trend line if data points above the red arrow and in the circle are ignored.

Figure 58 Variability of transmissivity in relation to distance away from fault considered at depth of borehole
end. Black line shows a possible trend line if data points above the red arrow and in the circle are ignored.

Page | 69

Distances at depth between boreholes and faults are worked out by using simple
geometry. The data considered is the topographical level, the borehole depth, the
dip direction of the particular fault being either in the direction or away from the
borehole and by assuming an average fault dip angle of 65o which is approximately
equal to an average fault dip (Appendix I). An interesting observation is the less
scatter of the data shown by the trend line for the plot considering the distance to
the faults at depth (Figure 58). Even though the resolution of the data is low, this
result is encouraging for further investigation on the level of control by faults on the
hydrogeology.

Page | 70

4.6

Potentiometry

Victoria fault

3km
Il-Maghlaq fault
Figure 59 1990 potentiometric map superimposed on the Geological Map of Malta (1993). Dashed lines show
main faults average alignments, dots with number show locations of gauged boreholes with water piezometric
level. (adapted from BRGM, 1991c & the Geological Map of Malta, 1993)

The readings of the period 1988 to 1991 are from 40 gauging boreholes which cover
a wide area of Malta (Figure 59). Although this number is not small it is surely not
enough to assess any fault control over water heads. Piezometers close to IlMaghlaq fault show the highest levels for readings in proximity to the coast. This
can be due to smeared clays which is known to occur at il-Maghlaq fault (Bonson et
al., 2007), however one should also be aware that the transmissivity for the LCLs at
depths greater than 100 metres below its top are reported to be low (BRGM,
1991b). This may be the case at this location since the top of the LCL here is about
100 metres above sea level (Pedley, 1975).

Page | 71

Discussion

The main scope of this chapter is to provide a discussion in a wider context of the
main knowledge acquired from the previous chapters especially from chapter 4. In
so doing the aims reached together with the limitations and further implications are
highlighted. These are then used to recommend future studies.

5.1

Geomorphologic Site Reconnaissance

The quantity of data that is acquired through this simple technique is appreciated.
Even though the extent of this data may not be detailed it provides a good basis as a
qualitative tool which guides further detailed field work and invasive investigation
methods.

5.1.1.

Stratal Dip of BC

It is understood that flow in the perched water aquifer of the BC is governed by the
down-dips of the BC stratum. Given that the strata are generally sub-horizontal
large thicknesses of this aquifer are not expected however slightly thicker localised
perched aquifers may occur at the fault drags caused by the ENE-WSW faulting.
These may be spatially limited, not least by a general dip to the east.
This knowledge may be applied to quantify the effects of flow on rock toppling
phenomena (Figure 16; Figure 25) and denser vegetation bands (Figure 21). It is
known that UCL rock toppling failure on top of the BC is triggered by the different
mechanical and hydrogeologic properties of these materials (Gianfranco et al., 2003;
Magri et al., 2008; Devoto et al., 2012).
Further implications to the latter statement include the characterisation of
geomorphologic variability and its correlation with other parameters so as to

Page | 72

characterise these slope instabilities. Possible studies might include the temporal
development of instabilities due to wetting and drying cycles and the combined
effect of discontinuity patterns. The study of desiccation cracks noted at l-Imgiebah
may be included in this study. What happens if the BC is submerged under sea
water? What is the difference from outcropping BC slopes?

5.1.2.

Flow indications from karst erosion

It has been observed that karst morphology is a function of grain size of the strata
and their fracturing, with larger developments of karst in coarse-grained strata and
highly fractured zones. Higher developments of karst in coarse-grained strata may
be due to inferior packing of particles and a better connectivity of voids. Other
variables such as flow quantities, the chemical composition such as the calcium
carbonate saturation and the temporal framework of this process are not studied.
In general relatively larger developed karst features are observed in the UCL and
LCL which are coarser-grained (Qammiegh, L-Imgiebah Bay, Fomm ir-Rih Bay, Wied
il-Ghasel, Gharghur and St. Georges Bay). The development of karst within the GL
which are finer-grained is limited in width and generally follows only one direction
along the main wider discontinuities karst features are observed (Qammiegh, LImgiebah, Fomm ir-Rih Bay, Msida, Xghajra, Munxar and Birkirkara).
Karstified caves are noted at a fault zone near Qammiegh (Figure 18) while at
Gharghur karstified notches in the cliff face are noted to develop from a wider
aperture (Figure 34).
Several observations show that fluid conducting boundaries can be found between
layers of different grain size distribution. Perhaps the clearest evidence of this is the
karstified eroded boundary between the GL and LCL observed at Xghajra (Figure 47).
In addition observations of karst features and a dense band of vegetation at
Gharghur close to contacts of members from the same formation indicate that
these fluid conducting boundaries exist even between members or facies within the
Page | 73

same formation (Figure 33). Variation of grain size distributions therefore provides a
barrier to flow into the finer-grained materials at contacts between coarse-grained
and fine-grained materials and encourage flow along these boundaries.
A detailed geomorphologic exercise might identify more of these boundaries and
potentially aid in devising an investigation program to quantify this variability.
Ideally a test programme should aim at obtaining hydraulic data of the different
formations and facies. For Malta this data is widely missing. Further implications of
such an exercise may be the characterisation of geotechnical behaviour of geologic
materials in terms of their nature and state. These properties aid detailed
geotechnical design as for example in quantifying effective stresses.

5.1.3.

Sedimentation processes

Sedimentation processes are responsible for various material properties such as


mineralogy, grain size distribution and discontinuities.
In the MGL sub-horizontal discontinuities are interpreted as desiccation
discontinuities along sedimentation boundaries (Figure

30;

Figure 49). A study of

sedimentation environments and history may provide more information on more


possible boundaries.

5.1.4.

Calcite Deposition

Contrasting observations of karst features and calcite deposition onto rock surfaces
(Figure 35) is evidence of two contrasting affects water can have on carbonates.
This depends on the level of carbonate saturation of the water with respect to the
carbonate content of the particular rock (Fetter, 2001).

Page | 74

The implications are that water may either act to increase or decrease the
permeability by solution processes or calcite deposition respectively. The
characterisation of these processes requires the use of geochemistry.

5.1.5.

Style of faulting

St. Georges Bay provides hints on possible joint and fault styles (Figure 37; Figure
38; Figure 39). Hydraulic conductivity depends on the type of discontinuity
developed. Lower hydraulic conductivity is expected for infilled joints and higher
hydraulic conductivity expected for fault zone occupied by fault blocks and open
joints. Fault breccias may reduce the hydraulic conductivity of a fault. The
characterisation of faulting styles and parameters controlling their style are desired.
An antithetic fault is observed at Fomm ir-Rih Bay (Figure 28). Trends of
discontinuities with similar strikes but opposite dip directions are identified in some
of the sites (Appendix E). This may be linked with either faults bounding basins that
are widely observed in the horst and graben of North Malta or may be a matter of
mechanical variabilities of strata and thus their different mode of fracturing (Michie
et al., 2014).

5.2

Discontinuity data
5.2.1.

Jointing link with the rifting tectonics of Malta

It is observed that strikes and dip angles of the most occurring joints closely
resemble Maltas two main fault families being the ENE-WSW and NW-SE trending
faults. The most often identified joint sets are J1 and J2 (ENE-WSW trending)
followed by J3 (NW-SE trend). J1 and J2 are identified at four out of five sites while
J3 is only clearly identified at two of the sites. J3 not being identified at Xghajra is
probably due to the scan line orientation bias as it is oriented close to the NW-SE
direction. At Msida the reason cannot be identified clearly as various orientations
Page | 75

were considered in taking discontinuity data and all visible discontinuities were
measured.
The probable joint origin being the rifting tectonics is also evident from aperture
data. Concentration of data within the wider apertures is observed for joint sets J1,
J2 and J4 which closely resemble an E-W orientation. Joint sets J6, J7 & J9 have
concentration of data within the smaller end of aperture width classes. Their
orientation is approximately NNE-SSW which is furthest away from the orientation
of the main fault families of Malta. This supports a hypothesis that wider apertures
are noted for joints with orientations similar to the orientations of the main faults
of Malta.
Antithetic joint sets appear as trends in various sites. This may be another close link
with the horst and graben structure of Malta.

5.2.2.

Limited data from persistence

Some observations from persistence are also possible however it is better not to do
generalizations from this data given the evident data bias from low height of rock
wall outcrops studied.

5.2.3.

The case of the Birkirkara site

The case of the Birkirkara site is somewhat different from all the other sites. This
site lies at a distance of about 3 kilometres away from main mapped faults (the
Geological Map of Malta, 1993) while all the other sites lie within 1.5 kilometres of
main faults. Only J8 was identified as a main trending joint set here. J8 has a strike
similar to that of J1 and J2, however its average dip at this site is of 25o which is
shallow when compared to both the averages of J1 and J2 (see Table 7 and Table 8).
J8 reflects a previous event which is reported as up-arching of this member that is
responsible for the thickening of the LGL (Pedley, 1975; Pedley et al., 1976).
Page | 76

5.2.4.

Extent of tectonic affect

For all the sites except for Birkirkara it is believed that the discontinuity patterns
identified are closely linked to the tectonic origin of the same main faults that
outcrop on Malta. This is believed to be true even at distances of about 1.5
kilometres from mapped faults. Putz-Perrier and Sanderson (2010) note differences
in the deformation characteristics between higher strain zones, i.e. where faults
develop, and low strain zones. Considering that high strain zones could be a few
kilometres wide may be a supporting argument for the hypothesis that most
jointing in the sites studied has a tectonic origin.
The degree of aperture, persistence and frequency depend on the distance away
from faults. This is supported by higher relative K calculated for sites nearer to fault
zones such as Fomm ir-Rih Bay, Xghajra and St. Georges Bay while considerably
lower relative K at Msida. At Fomm ir-Rih Bay the highest relative K is for joint set J4
(ESE-WNW orientation) followed by joint set J3 (NW-SE orientation) representing
most strongly higher permeability in a direction sub-parallel to the orientation of
the nearest fault. At St. Georges Bay relative K of joint sets J1, J2, J3, J4 & J5 lie in
the same region. At the latter site the joint sets have a higher relative K resemble
closely the two main fault families of Malta and the bedding.
Further testing of the hypotheses presented in this section is required through a
larger sample size so as to correctly capture and represent all the variability that
may be at play. In such a study the effect from local geologic factors should be
identified.

5.2.5.

Control on hydraulic conductivities from geologic contacts

The highest relative K at Xghajra happens to be for joint set J10 which is a bedding
set and probably indicates the large effect on the hydraulic conductivity from the
GL-LCL contact.

Page | 77

5.2.6.

Jointing link with nearest fault structure

The ENE-WSW faults have varying strikes ranging from approximately E-W to NE-SW.
If considered at a local level, J1 and J2 closely follow the orientation of the nearest
fault orientation. This occurrence was noted at sites up to approximate distances of
1.5 kilometres away from a fault. However this is not the case for St. Georges Bay
at which site the orientation of J1 is closer to the orientation of the west Victoria
fault portion (Figure 60). A general fault strike kink in the Victoria fault is noted
which also coincides with an isopach thickening of the LGL south of the Victoria
fault at this location (Pedley, 1975). The average orientation of J1 at St. Georges
Bay may indicate that the Victoria fault might have had an effect on a wider zone
than the mapped fault as indicated by its hypothetical extension of the line but may
have been deviated due to some reason such as stress axis orientation or structural
thickening.
En-echelon left-stepping
east Victoria fault portion

St. Georges Bay

General fault trend kink


Fomm ir-Rih Bay

Average orientation of the


west Victoria fault portion

Hypothetical extension of
the orientation of the west
Victoria fault portion

3km
Figure 60 The Victoria fault within the regional geology (adapted from the Geological Map of Malta, 1993)

5.2.7.

Further data limitations

For the GL the bedding trend was in general not identified except for the Xghajra
site. A reason for this can be due to the GL being massively bedded (Pedley et al.,
1976) and thus not capturing the bedding trends by biased horizontal scan lines.

Page | 78

A wide dispersal of data is noted when measurements are in close proximity to


faults or other geologic structures. In such a case, in order to be able to reach a
regional comparison, probably a better approach would be to try and identify the
main geologic structures and hence the causes of the variability within the jointing.
When this cannot be achieved taking many readings as possible to try and obtain
the main trends is a good tool to eliminate as much as possible any data bias.
More trends of spatial variability of jointing within the different geological
formations and facies and within the different geologic contexts should be
identified. These should be backed by statistically significant samples. Ideally this
data should be correlated with field mapping and specifically designed testing
programmes.

5.3

Transmissivity

It is not usual to correlate transmissivity data to distance from faults. This was
carried out at the absence of any more detailed data. Even though a wide dispersal
of data is noted, a certain degree of link between the two seems evident (Figure 57,
Figure 58). At this stage this is an encouraging result considering the wide spatial
area represented by these tests and thus the variability they may represent. Ideally
specific data such as investigation borehole data is to be available so as to be able
to identify possible correlations of data. Such data is probably available at the hands
of the Maltese authorities however no such data is available to this author at the
time of writing this study.
It should also be noted that inferred faults marked as dashed lines of the Geological
Map of Malta (1993) are within a 50 metre accuracy bracket (Pedley, 2014).

Page | 79

5.4

Potentiometry

The potentiometric data available is widely spaced over most of Malta with no
concentration of closely installed piezometers. This does not allow proper analysis
of fault control on the hydraulic properties of the geology of Malta. Given that the
highest reported total heads are in the region between 3 to 5 metres (BRGM, 1991c;
Sapiano et al., 2006) the fault seal control on the hydrogeology is not expected to
be significant however heads depend also on the saltwater-freshwater contact.

5.5

Main Data Limitations

The methods applied and analyses carried out with the data at hand are in no way
exhaustive to understand the full extent of the hydrogeology of Malta. They give a
good indication of what the main controls on hydrogeology are and thus guide
further work.
A good number of sites were visited during the field trip gathering a good spatial
representativeness. Still, most probably they do not encompass the whole
variability of the hydrogeology of Malta. A lot of observations are made, however in
order to support the interpretations a higher number of similar observations across
more sites is desired. Some level of quantification or testing should be done. As a
start, discontinuity data was collected from across five sites.
Confidence in the accuracy of dip angles and dip directions is high. Confidence is
slightly lower for the other discontinuity characteristics since they were visually
taken. However it is still believed that they have a good level of accuracy. As regards
the scan line sample sizes and orientations it was shown that maybe longer scan
lines and more varied orientations are desirable. An improvement on the
discontinuity data set is to include more sites. If a fault such as Victoria fault needs
to be studied in more detail perhaps scan lines along and across the fault at chosen
sites fulfil this requirement. Techniques such as circular scan lines to measure

Page | 80

intensity, density and average length of discontinuities may capture data that is not
captured by linear scan lines.
The major limitation with the analysis carried out with the pre-available data is that
in general the data used may lack specific or complete detail. In the earth sciences
and the geotechnical field a lot of variables and unknowns are involved, so generally
several points of perspectives are desired so as to confirm or otherwise the data.
Possibilities of doing this with the available data are limited. In addition using the
Geological Map of Malta (1993) to measure distances to faults and topographical
levels is expected to result in low confidence outputs.

5.6

Conceptual Ground Model

The area shown is the area chosen for the conceptual ground model presented in
this section (Figure 61). The conceptual ground model (Figure 62) summarises a
good number of observations highlighted in this study.

N
Fomm ir-Rih Bay

2km

Figure 61 Red border shows the area of the conceptual ground model presented in this section (adapted from
the Geological Map of Malta, 1993)

Page | 81

Figure 62 Conceptual Ground Model highlighting regional hydrogeology of Malta. Annotations crossreferenced to numbers are included on the next page.

Page | 82

Displacements of faults towards the east coast probably decrease; therefore


the related fault damage zones narrow down. The synclinal structure due to
the fault drag also narrows down. As displacements decrease towards the east
a wedge-shaped block dipping towards the west is created.

Wider fault damage zones to the west of the Victoria Fault.

At larger displacements a certain width of damage zones are expected at the


LCL too but still considerably lower than those in the GL (refer to Table 12). For
small throws no damage width is expected within the LCL.

Widening of fault damage zones at GL probably increases slip surfaces.

BC smear provides a localized seal and compartmentalizes flow parallel to its


direction.

Confined Aquifer Zone? BC bottom dipping to a level below sea level

Fluid conduit GL-LCL contact. Is this conduit best developed at the saline-fresh
water contact? If yes do we expect a less extensive development of this
contact below sea-level?

No seal in NW-SE direction. Il-Maghlaq fault is absent from the coast closest to
this site. Therefore the potential of global aquifer blocks to be sealed by faults
is low. Transmissivity may vary due to variability in facies permeability.

Development of karst at GL along fault and subsequently along the subhorizontal GL-LCL contact?

10

Is BC seal breached for throws between 5-50 metres as observed by Missenard


et al. (2014)? Karstification of the underlying layers may depend on this.

11

Higher hydraulic conductivity at fault zones and thus lower total heads.

12

Gradual decrease of conductivity away from fault zone as joints apertures


close, frequency decreases and probably persistence decreases too. Higher
total heads (water table) are expected in these zones.

13

More developed karst may form at intersections between faults and sea levels
where highest levels of flow are expected. Vertical infiltration may be
inhibited by the clay cover.

14

The general dip of this wedge to the west is governed by the displacements
being largest to the west. Greater downthrows occur towards the west.

15

Valleys are formed dipping to the east in an approximate west-east direction.


Their occurrence is governed by the general dip of strata to the east and
probably by the ENE-WSW trending faults. Solution subsidence structures may
aid the development of valleys.

16

Springs

17

Sandy contact between the ULC and BC.

Page | 83

6
6.1

Summary and Conclusions


Main Conclusions

The main findings of this research work on the structural controls on the
hydrogeology of Malta are summarised.
The potentiometric data available is widely spaced over most of Malta with no
concentration of closely installed piezometers. This makes proper analysis of fault
seal control on the hydrogeology of Malta impossible. Evidenced in part by low
highest heads of about 3 to 5 metres, large aquifer blocks controlled by fault seals
are probably inexistent. Fault seals may be existent along a limited uni-directional
stretch.
Possibilities of fault parameters such as displacements to define the fault
architecture were previously shown (Michie et al., 2014). Hydraulic conductivity is
expected to vary depending on the type of discontinuity developed whether being
an infilled joint, made up of fault blocks and open joints or fault breccias.
In general the joints are shown to be closely linked to the latest rift tectonics of
Malta. This is evidenced from similar strikes and dip angles of the most occurring
joints that closely resemble the ENE-WSW and NW-SE trending faults, the wider
apertures of these joints and the relative K calculated. Even joints at distances of
about 1.5 kilometres away from mapped faults are shown to be linked with the rift
tectonics as their orientations closely resemble the nearest fault structures. A
decrease in aperture width and spacing is observed at locations away from the
faults. Confirmation or otherwise of this hypothesis should be tackled by collecting
a statistically significant sample.
The site of Birkirkara presents a somewhat different situation from all the other
sites. This site lies at a distance of about 3 kilometres away from the nearest
mapped faults (the Geological Map of Malta, 1993). The main trending joint set
identified, J8, shows strike similarity to the ENE-WSW trending faults but occurs at a
Page | 84

much shallower dip. This may show that rift tectonics do not have or have less of an
affect at these distances away and/or shows evidence of previous events that may
have been up-arching effects that are responsible for the isopach thickening of the
LGL (Pedley, 1975; Pedley et al., 1976).
From observations it is noted that karst development differs between formations or
facies that exhibit variability in grain size distribution and fracturing. Perhaps more
specifically coarse-grained strata may have inferior packing of particles and a better
connectivity of voids. Both of these two characteristics increase the hydraulic
conductivity of a rock mass. Therefore larger developments of karst are expected in
rock masses of higher hydraulic conductivities. In addition observations highlight
that fluid conducting boundaries can result between layers of different grain size
distributions. An erosive contact of the GL-LCL boundary at Xghajra is observed.
Sedimentation processes may be responsible for various material properties such as
grain-size distributions and discontinuities.
It is not usual to correlate transmissivity data to distance from faults especially
when the data is of low resolution. However even though a wide dispersal of data is
noted, a certain degree of correlation between the two seems plausible. This result
provides encouragement for future research.

6.2

Further Studies
6.2.1.

Fault parameters and control

The further study of fault architecture is encouraged along and across faults by
carrying out field mappings. Data gaps still exist by missing inland data. Correlations
from findings may shed light on hydraulic characterisation of faults. Techniques
such as linear and circular scan lines are deemed to aid the initial quantification.
A next step would be to design a field and laboratory testing programme so as to
correlate hydraulic properties to the geologic contexts identified. Methods to
Page | 85

correlate investigation data at depth with surface data are to be found. In so doing
data may be extrapolated to predict hydraulic behaviour at sites where data is still
limited. This data would then be subject to further future testing.

6.2.2.

Controls on jointing

More trends of spatial variability of jointing within the different geological


formations and the different geologic contexts should be tested by statistically
significant samples. Sample sites should preferably be chosen with the hydrological
cycle in mind.

6.2.3.

Permeability variability of different formations and facies

In this study controls on fluid flow paths due to formations or facies having varying
grain size distributions of the geological strata has been identified. The zoning of
such strata combined by intact rock and rock mass permeability testing would
provide useful information for characterising the hydrogeology of Malta. For Malta
this data is widely missing.
If this data is combined with geotechnical properties of geologic materials both in
terms of their nature and state useful information in quantifying effective stresses
for engineering purposes may be reached.

6.2.4.

Geochemistry

Both observations of carbonate solution and calcite deposition have been observed.
The geochemical processes at play have a governing effect on the hydraulic
properties of such zones. Analysis of flow paths from geomorphology in
combination with geochemical testing are deemed to provide a useful tool in
characterising possible geochemical controls on the hydrogeology of Malta.
Page | 86

References
Alexander, D. (1988) A review of the physical geography of malta and its significance
for tectonic geomorphology. Quaternary Science Reviews. 7 (1), 41-53.
Argnani, A. (1990) The strait of sicily rift zone: Foreland deformation related to the
evolution of a back-arc basin. Journal of Geodynamics. 12 (24), 311-331.
Anzidei, M., Baldi, P., Casula, G., Galvani, A., Mantovani, E., Pesci, A., Riguzzi, F. &
Serpelloni, E. (2001) Insights into present-day crustal motion in the central
Mediterranean area from GPS surveys. Geophys. J. Int. 146, 98-110.
ATIGA. (1972) Wastes disposal and water supply project in Malta. Malta, UNDP.
Bakalowicz, M. & Mangion, J. (2003) The limestone aquifers of Malta: Their
recharge conditions from isotope and chemical surveys. IAHS-AISH Publication.
(278), 49-54.
Bonson, C. G., Childs, C., Walsh, J. J., Schpfer, M. P. J. & Carboni, V. (2007)
Geometric and kinematic controls on the internal structure of a large normal fault in
massive limestones: The Maghlaq Fault, Malta. Journal of Structural Geology. 29 (2),
336-354.
Bosence, D. W. J. & Pedley, H. M. (1982) Sedimentology and palaeoecology of a
Miocene coralline algal biostrome from the Maltese Islands. Palaeogeography,
Palaeoclimatology, Palaeoecology. 38 (12), 9-43.
Bureau de Recherche Gologique et Minire (BRGM). (1991) Study of the freshwater resources of Malta. Executive Summary. Malta, Government of Malta.
Bureau de Recherche Gologique et Minire (BRGM). (1991b) Study of the freshwater resources of Malta. Appendix 2. Hydrodynamic characteristics. Malta,
Government of Malta.

Page | 87

Bureau de Recherche Gologique et Minire (BRGM). (1991c) Study of the freshwater resources of Malta. Appendix 4. Potentiometry of the mean sea level aquifer.
Malta, Government of Malta.
British Standards Institution (1999) BS 5930:1999. Code of practice for site
investigations. London, BSI.
Cartwright, J. A., Trudgill, B. D. & Mansfield, C. S. (1995) Fault growth by segment
linkage: an explanation for scatter in maximum displacement and trace length data
from the Canyonlands Grabens of SE Utah. Journal of Structural Geology. 17 (9),
1319-1326.
Debono, G. & Xerri, S. (1993) Geological Map of the Maltese Islands. 1:25000. Sheet
1 - Malta. Malta, Oil Exploration Directorate, Office of the Prime Minister.
Dart, C. J., Bosence, D. W. J. & McClay, K.R. (1993) Stratigraphy and structure of the
Maltese graben system. Journal of the Geological Society. 150, 1153-1166.
Dawers, N. H. & Anders, M. H. (1995) Displacement-length scaling and fault linkage.
Journal of Structural Geology. 17 (5), 607-614.
Devoto, S., Biolchi, S., Bruschi, V., M., Furlani, S., Mantovani, M., Piacentini, D.,
Pasuto, A. & Soldati, M. (2012) Geomorphological map of the NW Coast of the
Island of Malta (Mediterranean Sea). Journal of Maps. 8 (1), 33-40.
Faerseth, R. B. (2006) Shale smear along large faults: continuity of smear and the
fault seal capacity. Journal of the Geological Society, London. 163, 741-751.
Fetter, C. W. (2001) Applied Hydrogeology. Fourth Edition. New Jersey, USA,
Prentice-Hall, Inc.
Gardiner, W., Grasso, M. & Sedgeley, D. (1995) Plio-pleistocene fault movement as
evidence for mega-block kinematics within the HybleanMalta Plateau, Central
Mediterranean. Journal of Geodynamics. 19 (1), 35-51.

Page | 88

Gianfranco, M., Samori, L., Ragazzini, A., Cuppini, G. & Baratin, L. (2003)
Consolidamento del Palazzo Vilhena - M'dina (Malta) - Relazione geotecnica.
Bologna, Universita` di Bologna. Report number: 6.
Gillespie, P. A., Walsh, J. J. & Watterson, J. (1992) Limitations of dimension and
displacement data from single faults and the consequences for data analysis and
interpretation. Journal of Structural Geology. 14 (10), 1157-1172.
Gonzalez de Vallejo, L. I. & Ferrer, M. (2011) Geological Engineering. London, CRC
Press.
Google. (2014) Google Earth [computer programme] Google Inc.
Illies, J. H. (1981) Graben Formation - The Maltese Islands - a case history.
Tectonophysics. 73, 151-168.
Jongsma, D., Woodside, J. M., King, G. C. P. & van Hinte, J. E. (1987) The Medina
Wrench: a key to the kinematics of the central and eastern Mediterranean over the
past 5 Ma. Earth and Planetary Science Letters. 82 (12), 87-106.
Kim, Y., Peacock, D. C. P. & Sanderson, D. J. (2003) Mesoscale strike-slip faults and
damage zones at Marsalforn, Gozo Island, Malta. Journal of Structural Geology. 25
(5), 793-812.
Magri, O., Mantovani, M., Pasuto, A. & Soldati, M. (2008) Geomorphological
investigation and monitoring of lateral spreading along the north-west coast of
Malta. Geogr. Fis. Dinam. Quat. 31, 171-180.
Marrett, R. & Allmendinger, R. W. (1990) Kinematic analysis of fault-slip data.
Journal of Structural Geology. 12 (8), 973-986.
Ministry for European Affairs (2014) CF120: National Flood Relief Project (NFRP).
[Online]

Available

from:

https://investinginyourfuture.gov.mt/project/waste-

Page | 89

management-and-risk-prevention/national-flood-relief-project-nfrp-42041344
[Accessed 11th August 2014].
Mylroie, J. R. & Mylroie, J. E. (2007) Development of the carbonate island karst
model. Journal of Cave and Karst Studies. 69 (1), 59-75.
Micarelli, L., Benedicto, A. & Wibberley, C. A. J. (2006) Structural evolution and
permeability of normal fault zones in highly porous carbonate rocks. Journal of
Structural Geology. 28 (7), 1214-1227.
Michie, E. A. H., Haines, T. J., Healy, D., Neilson, J. E., Timms, N. E. & Wibberley, C. A.
J. (2014) Influence of carbonate facies on fault zone architecture. Journal of
Structural Geology. 65 (0), 82-99.
Missenard, Y., Bertrand, A., Vergly, P., Benedicto, A., Cushing, M. & Rocher, M.
(2014) Fracture-fluid relationships: implications for the sealing capacity of clay
layersInsights from field study of the Blue Clay formation, Maltese islands. Bulletin
De La Societe Geologique De France. 185 (1), 51-63.
Newbery, J. (1968) The perched water table in the upper limestone aquifer of
Malta. Journal of the Institution of Water. 22, 551-570.
Newbery, J. (1976) Miocene sea-floor subsidence and later subaerial solution
subsidence structures in the Maltese Islands. Proceedings of the Geologists'
Association. 87 (1), 111.
Pedley, H. M. (1975) The Oligo-Miocene sediments of the Maltese Islands. PhD. The
University of Hull.
Pedley, H. M. (1975b) Miocene Sea-floor Subsidence and Later Subaerial Solution
Subsidence Structures in the Maltese Islands. Proc. Geol. Ass. 85 (4), 533-547.

Page | 90

Pedley, H. M. (1987) Controls on Cenozoic carbonate deposition in the Maltese


Islands: review and reinterpretation. Memorie Della Societa Geologica Italiana. 38,
81-94.
Pedley, H. M. (1990) Syndepositional tectonics affecting Cenozoic and Mesozoic
deposition in the Malta and SE Sicily areas (Central Mediterranean) and their
bearing on Mesozoic reservoir development in the N Malta offshore region. Marine
and Petroleum Geology. 7 (2), 171-180.
Pedley, M. (1998) A review of sediment distributions and processes in OligoMiocene ramps of southern Italy and Malta (Mediterranean divide). Geological
Society, London, Special Publications. 149, 163-179.
Pedley, M. (2011) The Calabrian Stage, Pleistocene highstand in Malta: a new
marker for unravelling the Late Neogene and Quaternary history of the islands.
Journal of the Geological Society, London. 168, 913-925.
Pedley, Martyn. Geologist. (Personal communication, 6th August 2014).
Pedley, H. M., House, M. R. & Waugh, B. (1976) The geology of Malta and Gozo.
Proceedings of the Geologists' Association. 87 (3), 325-341.
Pedley, H. M. & Bennett, S. M. (1985) Phosphorites, hardgrounds and
syndepositional solution subsidence: A palaeoenvironmental model from the
miocene of the Maltese Islands. Sedimentary Geology. 45 (1-2), 1-34.
Pratt, S. K. (1990) Hardground genesis in pelagic carbonates from the Miocene of
Malta and Cretaceous of Southern England. PhD. University of London.
Putz-Perrier, M. W. (2008) Distribution and Scaling of Extensional Strain in
Sedimentary Rocks. PhD. Imperial College of Science, Technology and Medicine, UK.

Page | 91

Putz-Perrier, M. W. & Sanderson, D. J. (2010) Distribution of faults and extensional


strain in fractured carbonates of the North Malta Graben. The American Association
of Petroleum Geologists. 94 (4), 435-456.
Reuther, C. -D. & Eisbacher, G. H. (1985) Pantelleria Rift - crustal extension in a
convergent intraplate setting. Geologische Rundschau. 74 (3), 585-597.
Rocscience Inc. (2004), Dips Version 5.1 - Graphical and Statistical Analysis of
Orientation Data. www.rocscience.com, Toronto, Ontario, Canada. [Computer
program].
Sapiano, M., Mangion, J. and Batchelor, C. (2006) Water Resources Review. Rome,
FAO. Report number: 1.
Sapiano, Manuel. Head of water policy unit within MRA. (Personal communication,
31st July 2014).
Seebeck, H., Nicol, A., Walsh, J. J., Childs, C., Beetham R. D. & Pettinga, J. (2014)
Fluid flow in fault zones from an active rift. Journal of Structural Geology. 62, 52-64.
Snow, D. T. (1970) The frequency and apertures of fractures in rock. J. Rock Mech.
Min. Sci. 7, 23-40.
Stuart, M. E., Maurice, L., Heaton, T. H. E., Sapiano, M., Micallef Sultana, M.,
Gooddy, D. C. & Chilton, P. J. (2010) Groundwater residence time and movement in
the Maltese islands A geochemical approach. Applied Geochemistry. 25 (5), 609620.
Trechmann, C. T. (1938) Quaternary Conditions in Malta. The Geological Magazine.
75 (1), 1-26.
Watterson, J. (1986) Fault Dimensions, Displacements and Growth. Pageoph. 124,
365-373.

Page | 92

Appendix A The Geological Map of Malta (1993)

Page | 93

Page | 94

Appendix B Main water bodies as indicated by the Malta Resources


Authority (MRA)

Page | 95

Groundwater Body Code


MT001
Groundwater Body Name
Malta Main Mean Sea Level Groundwater Body
Reference Year
2004

General Characteristics
Location
The Malta Main Mean Sea Level Groundwater Body is sustained in the Lower Coralline Limestone aquifer and is in free contact with sea-water.
This groundwater body extends over the whole southern and central parts of the Island, under the Rabat Dingli Plateau, The Mgarr Plateau, the
Wardija Ridge up to the Pwales Valley as its northern boundary.
In real terms the Groundwater Body can be compared to a lens-shaped body of fresh-water floating on more saline water, with a thickness of
fresh-water below sea-level approximately thirty-six times its piezometric height above sea level.
Area

217km2

Main Aquifer

Lower Coralline Limestone

Main Aquifer Type

Fractured Carbonate Media

Groundwater Horizon

1; 2 in the western regions

Maximum Length

21km

Maximum Width

13km

Mathematical centre of groundwater body

450600, 3971400

Hydro-geological characteristics
Stratigraphy

TertiaryOligocene

Mean Annual Precipitation

543mm

Mean Groundwater Body Thickness

67.5m

Main Recharge Source

Precipitation

Mean Annual Recharge

34.3hm3

Pressures
Main Land-Use Features (Corinne Landcover 2000)
Discontinuous urban fabric

23%

Agriculture with significant area of natural vegetation

43%

Schlerophyllous vegetation

6%

Sparsely Vegetated areas

3%

Areas overlain by perched aquifers

17%

Industrial zones

4%

Airport

2%

Mineral abstraction sites

1%

Other Pressures
Water Abstraction Purpose

Potable Supply, Irrigation, Secondary Domestic and Industrial

Artificial Recharge

Mainly due to leakages from the potable supply and sewerage network

Contaminated Land

Old un-lined landfill sites at Maghtab, Wied Fulija and Luqa

Possible Associated Aquatic Ecosystems

Is-Salina, Il-Maghluq (Marsascala), Il-Ballut (Marsaxlokk) L-Ghadira


s-Safra

Page|97

Page|98

Page|99

Page|100

Page|101

Page|102

Page|103

Page|104

Appendix C Mdina investigation boreholes from Gianfranco et al.


(2003)

Page | 105

Page | 106

Page | 107

Page | 108

Page | 109

Page | 110

Page | 111

Page | 112

Page | 113

Page | 114

Page | 115

Appendix D Stereonet plots

Fomm ir-Rih Bay stereonet plot Both Scan Lines

Joint Set
1
3

Dip
70o
80o

Dip direction
N359o
N255o

Fomm ir-Rih Bay stereonet plot Scan Line A

Joint Set
1
3

Dip
89o
79o

Dip direction
N005o
N255o

Page | 116

Fomm ir-Rih Bay stereonet plot Scan Line B

Page | 117

St. Georges Bay stereonet plot All Scan Lines

Joint Set
1
3
5 (bedding)

Dip
86o
66o
11o

Dip direction
N356o
N259o
N057o

St. Georges Bay stereonet plot Scan Line A

Joint Set
1
3
5 (bedding)

Dip
86o
58o
12o

Dip direction
N356o
N237o
N060o

Page | 118

St. Georges Bay stereonet plot Scan Line B

Joint Set
1
2
3
6 (bedding??)

Dip
80o
84o
68o
9o

Dip direction
N353o
N192o
N258o
N314o

St. Georges Bay stereonet plot Scan Line C

Joint Set
2
3
4
5 (bedding)
7

Dip
68o
70o
89o
20o
74o

Dip direction
N158o
N217o
N031o
N045o
N117o

Page | 119

Msida stereonet plot All faces

Joint Set
1&2

Dip
90o

Dip direction
N337o /N157o

Msida stereonet plot Faces A, E and Face Parallel to A

Joint Set
1
7

Dip
84o
61o

Dip direction
N340o
N113o

Page | 120

Msida stereonet plot Face B

Joint Set
1&2

Dip
90o

Dip direction
N336o / N156o

Msida stereonet plot Face D

Page | 121

Xghajra stereonet plot Both Scan Lines

Joint Set
1
2
5 (bedding)

Dip
85o
85o
5o

Dip direction
N325o
N149o
N056o

Xghajra stereonet plot Scan Line A

Joint Set
2
5 (bedding ??)

Dip
87o
13o

Dip direction
N144o
N090o

Page | 122

Xghajra stereonet plot Scan Line B

Joint Set
1
2
5 (bedding)

Dip
83o
83o
5o

Dip direction
N324o
N152o
N045o

Page | 123

Birkirkara stereonet plot All Scan Lines

Joint Set
8

Dip
25o

Dip direction
N174o

Birkirkara stereonet plot Scan Line A

Joint Set
8

Dip
22o

Dip direction
N180o

Page | 124

Birkirkara stereonet plot Scan Line B

Joint Set
8

Dip
35o

Dip direction
N165o

Page | 125

Appendix E Full scan lines sheets

Page | 126

Fomm ir-Rih Discontinuity Scan Lines (letter in first column refers to scan line and number to position in metres)

Page | 127

St Georges Bay Discontinuity Scan Line A (letter in first column refers to scan line and number to position in metres)

Page | 128

Page | 129

Page | 130

Page | 131

St Georges Bay Discontinuity Scan Line B (letter in first column refers to scan line and number to position in metres)

Page | 132

Page | 133

Page | 134

St Georges Bay Discontinuity Scan Line C (letter in first column refers to scan line and number to position in metres)

Page | 135

Page | 136

Page | 137

Msida Discontinuity Faces A & E (letter in first column refers to face and number refers to reading entry)

Page | 138

Msida Discontinuity Faces B & C (letter in first column refers to face and number refers to reading entry)

Page | 139

Msida Discontinuity Face D (letter in first column refers to face and number refers to reading entry)

Page | 140

Xghajra Discontinuity Scan Line A (letter in first column refers to scan line and number to position in metres)

Page | 141

Xghajra Discontinuity Scan Line B (letter in first column refers to scan line and number to position in metres)

Page | 142

Birkirkara Discontinuity Face A (letter in first column refers to face and number refers to reading entry)

Page | 143

Birkirkara Discontinuity Face B (letter in first column refers to face and number refers to reading entry)

Page | 144

Appendix F Tables and graphs of discontinuities aperture data

Fomm ir-Rih Bay Aperture class frequencies for each joint set
Joint Set
/
Aperture
1

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

1
2

7
8
Totals

Fomm ir-Rih Bay Aperture class percentages for each joint set
Joint Set
/
Aperture

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

1
2

33.3

33.3

66.7

33.3

50
40

100

33.3

50

60

100

100

7
8
Totals

100

100

100

Page | 145

Page | 146

St George Bay Aperture class frequencies for each joint set


Joint Set
/
Aperture
1

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

11

11

14

13

11

Totals

35

26

32

38

37

16

14

J9

J10

St George Bay Aperture class percentages for each joint set


Joint Set
/
Aperture

J1

14.3

17.1

3.8

21.9

5.7

15.4

31.4

J2

J3

J4

J5

J6

J7

J8

10.8

6.3

7.1

11.1

21.1

24.3

6.3

35.7

11.1

22.2

9.4

7.9

10.8

18.8

22.2

11.1

33.3

42.3

43.8

34.2

16.2

25.0

35.7

44.4

55.6

16.7

14.3

15.4

18.8

18.4

29.7

43.8

21.4

11.1

11.1

33.3

5.7

3.8

13.2

5.4

5.7

11.5

3.1

5.3

2.7

5.7

7.7

Totals

100

100

100

100

100

3.1

16.7

100

100

100

100

100

Page | 147

Page | 148

Msida Aperture class frequencies for each joint set


Joint Set
/
Aperture
1

J1

J2

J3

J4

4
5

J6

J7

J8

J9

J10

1
1

6
7

J5

1
2

14

15

8
Totals

Msida Aperture class percentages for each joint set


Joint Set
/
Aperture

J1

J2

J3

J4

28.6

53.3

50

66.7

40

50

28.6

26.7

50

16.7

60

50

21.4

100

100

4
5

J6

J7

J8

J9

J10

6.7
7.1

6.7

6
7

J5

16.7
14.3

6.7

100

100

8
Totals

100

100

Page | 149

Page | 150

Xghajra Aperture class frequencies for each joint set


Joint Set
/
Aperture
1

J1

J2

J3

J4

J5

J6

J7

2
3

J9

J8

J10

1
2

J8

J9

J10

50

33.3

7
8
Totals

Xghajra Aperture class percentages for each joint set


Joint Set
/
Aperture

J1

J2

J3

J4

J5

J6

J7

1
2
3

50

25

50

25

33.3
100

100

100

33.3
33.3

50

33.3

33.3

50

100

100

100

7
8
Totals

100

100

100

100

100

100

100

Page | 151

Page | 152

Birkirkara - Aperture class frequencies for each joint set


Joint Set
/
Aperture
1
2

J1

J2

J3

J4

J5

J6

J7

J9

J10

2
1

1
3

8
Totals

J8

3
2

1
4

10

J8

J9

J10

Birkirkara - Aperture class percentages for each joint set


Joint Set
/
Aperture

J1

J2

J3

J4

J5

J6

J7

1
2

25

10

28.6

25

25

25

57.1

100

20
50

100

40

100

14.3
42.9

14.3

8
Totals

28.6

12.5

37.5
25

50

30

25

14.3
100

100

100

100

100

100

100

100

100

Page | 153

Page | 154

Page | 155

Page | 156

Page | 157

Page | 158

Page | 159

Page | 160

Page | 161

Appendix G Tables and graphs of discontinuities persistence data

Fomm ir-Rih Bay Persistence class frequencies for each joint set
Joint Set
/
Persistence
1
2

J1

J3

J4

J5

J6

J7

J8

J9

J10

1
2

3
4

J2

5
Totals

Fomm ir-Rih Bay Persistence class percentages for each joint set
Joint Set
/
Persistence

J1

J2

J3

J4

1
2

J6

J7

J8

J9

J10

100
40

3
4

J5

66.7

50

33.3

25

60.0

66.7

25

33.3

100

100

5
Totals

100

100

100

Page | 162

& J4

& J4

Page | 163

St. Georges Bay Persistence class frequencies for each joint set
Joint Set
/
Persistence
1

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

15

11

20

13

12

17

10

10

22

17

4
5
Totals

1
36

26

32

38

37

16

14

St. Georges Bay Persistence class percentages for each joint set
Joint Set
/
Persistence

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

41.7

42.3

62.5

34.2

32.4

12.5

35.7

44.4

66.7

33.3

47.2

38.5

31.3

57.9

45.9

43.8

64.3

44.4

33.3

50.0

11.1

19.2

6.3

5.3

13.5

43.8

2.6

5.4

4
5
Totals

11.1

16.7

2.7
100

100

100

100

100

100

100

100

100

100

Page | 164

Page | 165

Msida Persistence class frequencies for each joint set


Joint Set
/
Persistence
1

J1

J2

J3

J4

10

Totals

10

J5

J6

J7

J8

J9

J10

2
2
2

15

Msida Persistence class percentages for each joint set


Joint Set
/
Persistence

J1

10

10

26.7

40

66.7

20

20

6.7

20

20

Totals

100

J2

J3

J4

50

40

50

20

J5

J6

J7

J8

J9

J10

50
100

50
100

100

100

100

100

Page | 166

Page | 167

Xghajra Persistence class frequencies for each joint set


Joint Set
/
Persistence
1

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

5
Totals

Xghajra Persistence class percentages for each joint set


Joint Set
/
Persistence

J1

J2

J3

J4

J5

50

50

50

100

100

50

100

J6

J7

J8

J9

J10

33.3

33.3

33.3

66.7

100

100

100

100

33.3

5
Totals

100

100

100

100

100

100

100

Page | 168

Page | 169

Birkirkara - Persistence class frequencies for each joint set


Joint Set
/
Persistence
1

J1

J2

J3

J4

J5

J6

J7

J8

J9

J10

4
5
Totals

1
4

2
2

10

Birkirkara - Persistence class percentages for each joint set


Joint Set
/
Persistence

J1

J2

1
2

25

37.5

75

50

J3

J4

50

62.5

50

J5

J6

J7

J8

100

J9

J10

100

25

12.5

50

20

50

25

50

60

25

4
5
Totals

12.5
100

100

20
100

100

100

100

100

100

100

Page | 170

Page | 171

Appendix H Relative hydraulic conductivities

Fomm ir-Rih Bay - Relative Hydraulic Conductivity for each joint set
Joint Number of
Set readings, n

Length of
scan line,
L (m)

Frequency,
i (m-1)

Average
Aperture,
ei (mm)

ei3
(mm3)

Relative
Ki
(mm3/m)

J1

14.1

0.43

1.5

3.375

1.436

J2

14.1

0.28

0.375

0.053

0.015

J3

14.1

0.28

8.9375

713.918

202.530

J4

14.1

0.43

17.5

5359.375

2280.585

14.1

0.07

1.5

3.375

0.239

J5
J6
J7
J8
J9
J10

St. George's Bay - Relative Hydraulic Conductivity for each joint set
Joint Number of
Set readings, n

Length of
scan line,
L (m)

Frequency,
i (m-1)

Average
Aperture,
ei (mm)

ei3
(mm3)

Relative
Ki
(mm3/m)

J1

36

68.3

0.53

1.5

3.375

1.779

J2

26

68.3

0.38

1.5

3.375

1.285

J3

32

68.3

0.47

1.5

3.375

1.581

J4

38

68.3

0.56

1.5

3.375

1.878

J5

37

68.3

0.54

1.5

3.375

1.828

J6

16

68.3

0.23

1.5

3.375

0.791

J7

14

68.3

0.20

1.5

3.375

0.692

J8

68.3

0.13

1.5

3.375

0.445

J9

68.3

0.13

1.5

3.375

0.445

J10

68.3

0.09

1.5

3.375

0.296

Page | 172

Msida - Relative Hydraulic Conductivity for each joint set


Joint Number of
Set readings, n

Length of
scan line,
L (m)

Frequency,
i (m-1)

Average
Aperture,
ei (mm)

ei3
(mm3)

Relative
Ki
(mm3/m)

J1

15

107.9

0.14

0.175

0.005

0.00075

J2

16

107.9

0.15

0.05

0.000

0.00002

J3

107.9

0.02

0.1125

0.001

0.00003

J4

107.9

0.06

0.05

0.000

0.00001

107.9

0.05

0.175

0.005

0.00025

107.9

0.02

0.1125

0.001

0.00003

J5
J6
J7
J8
J9
J10

Xghajra - Relative Hydraulic Conductivity for each joint set


Joint Number of
Set readings, n

Length of
scan line,
L (m)

Frequency,
i (m-1)

Average
Aperture,
ei (mm)

ei3
(mm3)

Relative
Ki
(mm3/m)

J1

29.5

0.24

0.375

0.053

0.013

J2

13

29.5

0.44

1.5

3.375

1.487

J3

29.5

0.03

6.25

244.141

8.276

J4

29.5

0.07

1.5

3.375

0.229

J5

29.5

0.20

1.5

3.375

0.686

J6

29.5

0.03

J7

29.5

0.14

6.25

244.141

33.104

J8

29.5

0.10

0.175

0.005

0.001

J9

29.5

0.10

0.375

0.053

0.005

J10

29.5

0.03

17.5

5359.375

181.674

Page | 173

Appendix I Hydrodynamic data

Page | 174

Compiled table used for the re-interpretation of Transmissivity (BRGM, 1991b; Sapiano, 2014)

Page | 175

Page | 176

Page | 177

Data points used shown on the Geologic Map of


Malta (1993).
First number is the ID number of the well
borehole and the second number is the
transmissivity in m2/s. Both values are
referenced with the tables presented.

Each square is a 1 x 1 km square.

Page | 178

Appendix J Fault data and structural contours of LCL

Page | 179

Fitting of data points to determine an equation in the form of


Section 3.6.1

In the following working

is represented by the letter

The general equation of a parabola is


constants such that

and

is represented by .
, where ,

and

. Hence substituting the three points

and

are
,

in the general equation, the following three equations

will be produced:
Equation (1):
Equation (2):
Equation (3):

Solving equations (2) and (3) simultaneously:


The following Equation (4) is obtained by multiplying Equation (2) by

Equation (4):
Equation (3):
Equation (4) Equation (3):

Substituting the value of

in Equation (3) we get

.
Page | 180

Hence, the equation is

The above quadratic equation is graphically shown as follows:

Research shows that the relationship between displacement ( ) and length of fault
( ) is of the form

, where

is a constant not equal to zero (Watterson,

1986; Cartwright et al., 1995). Hence, by adjusting the above quadratic equation by
varying the coefficient of

and fixing the coefficient of

approximation
of

and

equal to zero, we get the

. The negative sign of the coefficient

intercept are insignificant, as they only translate the graph on a

different position on the axes and do not alter the shape of the graph. Graphically
the adjusted equation is shown below.

Page | 181

Table showing fault data extracted from the Geological Map of Malta (1993)
Name of section
Linetrend

Section A-A` of The Geological Map of Malta


170
Dip
direction
o
[ from
N.]

deg. from North


Actual
Throw
[m]

App.
Heave
[m]

Corr.
Heave
[m]

Displacement
[m]

80.63

32.25

17.31

16.99

36.45

57.0

278.76

111.50

72.6

72.50

133.00

341

66.0

17.46

6.98

3.15

3.11

7.65

144

71.2

94.07

37.63

14.29

12.84

39.76

3,020.52

315

59.2

58.06

23.22

16.87

13.82

27.02

3,911.21

333

65.3

49.04

19.62

9.45

9.04

21.60

4,232.21

163

71.9

86.46

34.58

11.38

11.30

36.38

4,331.73

163

72.7

155.42

62.17

19.46

19.31

65.10

5,527.27

172

73.2

119.65

47.86

14.46

14.45

49.99

10

6,087.94

340

60.5

114.77

45.91

26.42

26.02

52.77

11

6,319.83

350

57.8

81.45

32.58

20.49

20.49

38.49

12

6,743.00

165

61.8

21.17

8.47

4.56

4.54

9.61

13

7,377.05

355

66.9

25.99

10.40

4.45

4.43

11.30

14

7,498.86

167

75.8

10.06

4.02

1.02

1.02

4.15

15

7,854.30

199

65.4

61.98

24.79

12.99

11.36

27.27

16

9,086.62

61.8

452.45

180.98

98.93

97.11

205.39

17

9,301.60

142

63.6

18.20

7.28

4.1

3.62

8.13

18

10,333.65

146

72.2

47.54

19.02

6.7

6.12

19.98

19

10,638.03

331

63.9

17.16

6.86

3.56

3.37

7.64

20

11,525.99

157

62.6

28.11

11.24

5.98

5.83

12.66

21

13,014.22

335

63.9

10.10

4.04

2.05

1.98

4.50

Actual
Dip
o
[]

Fault
Number

Distance
[m]

START

0.00

1,085.88

62.2

1,372.46

167

1,771.53

2,403.42

Measured
throw
[m]

Fault

Qammiegh

Victoria

Page | 182

Fault geometry (sourced from Michie et al., 2014) Section 3.6.2

Calculation of FSZ width using geometry from Michie et al. (2014)


a
b
x
wh
h
wf
FSZ

67 deg.
80 deg.
60.0
102.3
35.7
25.5
163.4

m
m
m
m
m

(average from Michie et al., 2014)


(average from Michie et al., 2014)
(GL thickness from Pedley, 1975)

However FSZ width is expected to be wider than this due to


presence of BC and UCL too at this point. So let us say not less than
200 metres.

Page | 183

Predictive equations for damage zones width against displacement (sourced from
Michie et al., 2014)

Page | 184

Calculations of damage zones widths using the predictive equations in the


previous page for Table 12 (sourced from Michie et al., 2014)

Heave
Fault dip
Displacement
Formation

10 m
74 deg.
36.3 m
GL

(avg. for GL and LCL formations - Michie et al., 2014)

LCL

FSZ (m)

22.2

2.1

TDZ (m)
Avg. Fault core (m)

71.2

14.3
0.6

15.3

(range for LCL between 2.1m


and 15.3m)

Page | 185

You might also like