You are on page 1of 77

Predicting Riverbank Collapse Using Numerical Modelling

Mark Jaksa, Yien Lik Kuo, Chen Liang, Bertram Ostendorf,


Tom Hubble and Elyssa De Carli

Goyder Institute for Water Research


Technical Report Series No. 15/41

www.goyderinstitute.org
Riverbank Collapse Task 5 Report

Goyder Institute for Water Research Technical Report Series ISSN: 1839-2725

The Goyder Institute for Water Research is a partnership between the South Australian Government
through the Department for Environment, Water and Natural Resources, CSIRO, Flinders University,
the University of Adelaide, the University of South Australia and ICE WaRM (the International Centre
of Excellence in Water Resources Management). The Institute will enhance the South Australian
Government’s capacity to develop and deliver science-based policy solutions in water management. It
brings together the best scientists and researchers across Australia to provide expert and
independent scientific advice to inform good government water policy and identify future threats and
opportunities to water security.

The following organisation contributed to this report:

Enquires should be addressed to: Goyder Institute for Water Research


Level 4, 33 King William St, Adelaide SA 5000
tel: (08) 8236 5209
e-mail: enquiries@goyderinstitute.org

Citation
Jaksa, MB, Kuo, YL, Liang, C, Ostendorf, B, Hubble, T, and de Carli, E. 2015, Predicting Riverbank Collapse Using
Numerical Modelling, Goyder Institute for Water Research Technical Report Series No. 15/41, Adelaide, South
Australia

Copyright
© 2015 University of Adelaide. To the extent permitted by law, all rights are reserved and no part of this
publication covered by copyright may be reproduced or copied in any form or by any means except with the
written permission of University of Adelaide.

Disclaimer
The Participants advise that the information contained in this publication comprises general statements based on
scientific research and does not warrant or represent the completeness of any information or material in this
publication.

Page 2 of 77
Riverbank Collapse Task 5 Report

Table of Contents

Executive Summary............................................................................................... 8

1 Introduction .....................................................................................................10

2 GIS-based Analysis of Riverbank Collapses at Long Island Marina and Tailem


Bend.................................................................................................................12
2.1 Introduction ............................................................................................................... 12
2.2 Background and Methodology .................................................................................. 14
2.3 Results ........................................................................................................................ 18
2.4 Summary .................................................................................................................... 19

3 Back Analysis of Historical Riverbank Collapses .................................................20


3.1 Introduction ............................................................................................................... 20
3.2 Methodology ............................................................................................................. 22
3.3 Results ........................................................................................................................ 29
3.4 Summary .................................................................................................................... 33

4 Effects of River Level Fluctuation and Climate on Riverbank Stability ................34


4.1 Background ................................................................................................................ 34
4.2 Study Area.................................................................................................................. 36
4.3 Methodology ............................................................................................................. 38
4.4 Results ........................................................................................................................ 47
4.5 Summary .................................................................................................................... 55

5 Identifying the Areas Susceptible to Riverbank Instability .................................56


5.1 Introduction ............................................................................................................... 56
5.2 Methodology and Results .......................................................................................... 56
5.3 Summary .................................................................................................................... 65

6 Summary ..........................................................................................................68

7 References .......................................................................................................69

Page 3 of 77
Riverbank Collapse Task 5 Report

List of Figures
FIGURE 1: LOCATION OF THE STUDY AREA. .......................................................................................... 12
FIGURE 2: LONG ISLAND MARINA STUDY SITE: (A) LOCATIONS OF 5 SIGNIFICANT FAILURES; (B)
LOCATION PLAN OF IN SITU TESTING AND RECORDED COLLAPSES; AND (C) DISTRIBUTION OF
BANK CROSS-SECTIONS IN 3D PERSPECTIVE. ................................................................................. 13
FIGURE 3: EXAMPLES OF VISUAL INTERPRETATION OF 2008 AND 2010 AERIAL IMAGES USING ARCGIS
AT: (A) MURRAY BRIDGE; AND (B) TAILEM BEND. ......................................................................... 13
FIGURE 4: RIVER MURRAY WATER LEVEL AT MURRAY BRIDGE 1/12/1986 TO 11/07/2011 (BASED ON
DFW, 2010). .................................................................................................................................... 15
FIGURE 5: SCHEMATIC OF THE LOCATIONS AND THE SELF-WEIGHT OF THE STRUCTURES ON THE
RIVERBANK. .................................................................................................................................... 16
FIGURE 6: MINIMUM FOS AND POTENTIAL SLIP SURFACE OF A DEEP-SEATED ROTATIONAL FAILURE
OBTAINED FROM SVSLOPE AT LOCATION NO. 21 WHEN THE WATER LEVEL WAS AT 0 M AHD. .. 16
FIGURE 7: BACK ANALYSES USING THREE GEOTECHNICAL MODELS. .................................................... 17
FIGURE 8: FACTORS OF SAFETY OF NEIGHBOURING CROSS SECTIONS (0 AND 0.5 M AHD). ................ 18
FIGURE 9: PREDICTIONS OF RIVERBANK SUSCEPTIBILITY WITH RIVER LEVELS AT (A) 0 M AHD AND (B)
0.5 M AHD. ..................................................................................................................................... 19
FIGURE 10: THE GEOGRAPHICAL LOCATIONS OF THE FOUR STUDIED SITES AND MAJOR RIVERBANK
COLLAPSES IN THE PAST, AS WELL AS THE PROXIMITIES OF THE BOREHOLE AND RIVER LEVEL
OBSERVATION STATIONS................................................................................................................ 21
FIGURE 11: ADOPTED VISUAL INTERPRETATION METHOD OF HIGH-RESOLUTION AERIAL IMAGES: (A),
(C), (E) AND (G) ARE AERIAL PHOTOGRAPHS ACQUIRED IN MARCH 2008 AT MN, WR, RFR AND
WS, RESPECTIVELY; (B), (D), (F), AND (H) ARE AERIAL PHOTOGRAPHS ACQUIRED IN MAY 2010 AT
MN, WR, RFR AND WS, RESPECTIVELY. .......................................................................................... 23
FIGURE 12: EXAMPLE OF ADOPTED ELEVATION COMPARISON METHOD ON DEMS AT WR (A) 1 M
RESOLUTION DEM ACQUIRED IN 2008; (B) 0.2 M RESOLUTION DEM ACQUIRED IN 2010). .......... 24
FIGURE 13: RESULTS OF CPTU SOUNDING (A) PROFILE; AND (B) PORE PRESSURE DISSIPATION TEST
RESULTS. ......................................................................................................................................... 27
FIGURE 14: DAILY RIVER LEVELS AND DAILY RAINFALL RECORDED AT (A) MANNUM (MN) SITE IN
APRIL 2009; (B) WOODLANE RESERVE (WR) SITE IN FEBRUARY 2009; (C) RIVER FRONT ROAD,
MURRAY BRIDGE (RFR) SITE IN FEBRUARY 2009; AND (D) WHITE SANDS (WS) SITE IN APRIL 2009.
........................................................................................................................................................ 28
FIGURE 15: RIVERBANK STABILITY ANALYSIS OF THE MANNUM (MN) SITE ON 21 APRIL 2009. .......... 30
FIGURE 16: RIVERBANK STABILITY ANALYSIS OF THE WOODLANE RESERVE (WR) SITE ON 26
FEBRUARY 2009. ............................................................................................................................. 30
FIGURE 17: RIVERBANK STABILITY ANALYSIS OF THE RIVER FRONT ROAD, MURRAY BRIDGE (RFR) SITE
ON 6 FEBRUARY 2009. .................................................................................................................... 30

Page 4 of 77
Riverbank Collapse Task 5 Report

FIGURE 18: RIVERBANK STABILITY ANALYSIS OF THE WHITE SANDS (WS) SITE ON 23 APRIL 2009. ..... 31
FIGURE 19: RIVERBANK COLLAPSE FACTOR OF SAFETY TIME SERIES FOR MANNUM (MN) IN APRIL
2009. ............................................................................................................................................... 31
FIGURE 20: RIVERBANK COLLAPSE FACTOR OF SAFETY TIME SERIES FOR WOODLANE RESERVE (WR) IN
FEBRUARY 2009. ............................................................................................................................. 32
FIGURE 21: RIVERBANK COLLAPSE FACTOR OF SAFETY TIME SERIES FOR RIVER FRONT ROAD (RFR) IN
FEBRUARY 2009. ............................................................................................................................. 32
FIGURE 22: RIVERBANK COLLAPSE FACTOR OF SAFETY TIME SERIES FOR WHITE SANDS (WS) IN APRIL
2009. ............................................................................................................................................... 32
FIGURE 23: DETAILS OF THE LONG ISLAND MARINA SITE. .................................................................... 37
FIGURE 24: RIVERBANK GEOMETRY DEFINITION. ................................................................................. 40
FIGURE 25: EXAMPLE OF ADOPTED VISUAL INTERPRETATION PROCESS ON HIGH-RESOLUTION, AERIAL
IMAGES WITHIN ARCGIS. ................................................................................................................ 41
FIGURE 26: GEOTECHNICAL PROFILES BASED ON SOIL SAMPLES TAKEN FROM SR-BH1 AND SR-CPTU6S
AT LONG ISLAND MARINA. ............................................................................................................. 42
FIGURE 27: PARTICLE SIZE DISTRIBUTIONS BASED ON THE SOIL SAMPLES FROM FOUR DIFFERENT
DEPTHS IN BOREHOLE SR-BH1........................................................................................................ 42
FIGURE 28: ESTIMATED SWCCS FOR THE THREE SOIL LAYERS AT LONG ISLAND MARINA USING THE
FREDLUND AND XING (1994) ESTIMATION METHOD. .................................................................... 43
FIGURE 29: DAILY RIVER LEVELS, DAILY RAINFALL AND DAILY MEAN TEMPERATURE FROM MAY 1,
2008 TO FEBRUARY 28, 2009 AT LONG ISLAND MARINA. .............................................................. 45
FIGURE 30: RESULTS OF (A) 2D (DAY 282: FEBRUARY 6, 2008) AND (B) 3D (DAY 287: FEBRUARY 10,
2008) RIVERBANK STABILITY ANALYSES OF LONG ISLAND MARINA SITE. ...................................... 46
FIGURE 31: EVOLUTION OF PORE WATER PRESSURE AT 5 SELECTED NODES THROUGH THE ENTIRE
RESEARCH PERIOD ACCOUNTING FOR, AND WITHOUT, EVAPORATION. ...................................... 48
FIGURE 32: PWP DISTRIBUTIONS AS A RESULT OF (A) THE HIGHEST (DAY 138) AND (B) LOWEST (DAY
302) RIVER LEVELS. ......................................................................................................................... 49
FIGURE 33: FACTORS OF SAFETY FROM THE 2D, 3D AND CRLM MODELS. ........................................... 50
FIGURE 34: FACTORS OF SAFETY FOR HISTORICAL MODEL (HM) AND CONSTANT RIVER STAGE MODEL
(CRLM) IN TWO SCENARIOS. .......................................................................................................... 53
FIGURE 35: MAGNIFIED RAINFALL MODEL (MRM) UNDER DIFFERENT RIVER LEVEL SCENARIOS. ....... 54
FIGURE 36: LOCATING DEEP HOLES NEAR THIELE RESERVE USING: (A) BATHYMETRY FROM HUBBLE
AND DE CARLI (2015) (AREAS OF BEDROCK MARGINS ARE REPRESENTED BY DASHED WHITE LINE);
(B) A CONTOUR MAP OBTAINED FROM ARCGIS SUPERIMPOSED ON SATELLITE IMAGERY FROM
GOOGLE MAPS; (C) BLUE CONTOUR LINES SHOWING ELEVATIONS THAT ARE –10 M AHD OR
DEEPER, WHILST THE GREEN CONTOUR LINES REPRESENT 0 M AHD. ........................................... 57

Page 5 of 77
Riverbank Collapse Task 5 Report

FIGURE 37: LOCATIONS OF DEEPENED CHANNEL BEDS DOWNSTREAM FROM BLANCHETOWN TO


WALKER FLAT (BLUE CONTOUR LINES SHOW ELEVATIONS THAT ARE AT –10 M AHD OR DEEPER,
WHILST THE GREEN CONTOUR LINES ARE AT 0 M AHD). ............................................................... 58
FIGURE 38: LOCATIONS OF DEEPENED CHANNEL BEDS DOWNSTREAM FROM WALKER FLAT TO
MURRAY BRIDGE EAST (BLUE CONTOUR LINES SHOW ELEVATIONS THAT ARE AT –10 M AHD OR
DEEPER, WHILST THE GREEN CONTOUR LINES ARE AT 0 M AHD). ................................................. 59
FIGURE 39: LOCATIONS OF DEEPENED CHANNEL BEDS DOWNSTREAM FROM MURRAY BRIDGE EAST
TO WELLINGTON (BLUE CONTOUR LINES SHOW ELEVATIONS THAT ARE AT –10 M AHD OR
DEEPER, WHILST THE GREEN CONTOUR LINES ARE AT 0 M AHD). ................................................. 60
FIGURE 40: A SERIES OF TRANSECTS ALONG 0 M AHD CONTOUR LINES NEAR THIELE RESERVE (BLUE
CONTOUR LINES SHOW ELEVATIONS THAT ARE AT –10 M AHD OR DEEPER, WHILST THE GREEN
CONTOUR LINES ARE AT 0 M AHD. THE RED DOT MARKS THE APPROXIMATE LOCATION OF THE
PAST RIVERBANK INSTABILITY INCIDENT AND THE DASHED ORANGE LINE MARKS THE AREAS OF
THE BEDROCK MARGINS). .............................................................................................................. 61
FIGURE 41: AN EXAMPLE OF AN SVSLOPE2D ANALYSIS (TRANSECT NO. 4212). ................................... 63
FIGURE 42: AN EXAMPLE OF THE RESULT OF AN SVSLOPE2D ANALYSIS (TRANSECT NO. 5529). ......... 63
FIGURE 43: RIVERBANKS THAT ARE VULNERABLE TO SLUMPING ARE IDENTIFIED (SHOWN IN YELLOW
DOTS) AFTER THE ANALYSES (ASSUMING A RIVER LEVEL AT 0 M AHD; BLUE CONTOUR LINES
SHOW ELEVATIONS THAT ARE AT –10 M AHD OR DEEPER WITH ONE METRE INTERVAL; RED DOT
SHOWS THE APPROXIMATE LOCATION OF THE OCCURRENCE OF PAST SLUMPING; DASHED
YELLOW LINES ARE AREAS OF BEDROCK MARGINS; AND LEVEES ARE REPRESENTED BY THE THICK
GREEN LINE). .................................................................................................................................. 64
FIGURE 44: INCLINATION OF THE RIVERBANKS NEAR THIELE RESERVE USING A 1 × 1 MATRIX OF
GRIDS. ............................................................................................................................................. 66
FIGURE 45: INCLINATION OF THE RIVERBANKS NEAR THIELE RESERVE USING A 10 × 10 MATRIX OF
GRIDS. ............................................................................................................................................. 67

Page 6 of 77
Riverbank Collapse Task 5 Report

List of Tables
TABLE 1: NATURE OF THE HISTORICAL RIVERBANK COLLAPSE-RELATED INCIDENTS AT 4 STUDY SITES.
........................................................................................................................................................ 20
TABLE 2: SOIL PROPERTIES FOR SATURATED AND UNSATURATED FLOW MODELLING. ....................... 26
TABLE 3: GEOTECHNICAL MODELS OF THE CLAY LAYER OBTAINED FROM BACK ANALYSES. ............... 29
TABLE 4: MODEL VALIDATION. .............................................................................................................. 31
TABLE 5: SOIL PARAMETERS FOR STABILITY ASSESSMENT. ................................................................... 41
TABLE 6: EQUATIONS USED TO CALCULATE FREDLUND AND XING (1994) SWCC FITTING PARAMETERS
BASED ON THE SOIL GRAIN SIZE DISTRIBUTION. ............................................................................ 44
TABLE 7: ADOPTED GEOTECHNICAL PARAMETERS FOR STABILITY CALCULATIONS. .............................. 62

Page 7 of 77
Riverbank Collapse Task 5 Report

Executive Summary
Unprecedented low river levels in the Lower River Murray (downstream of Lock 1, Blanchetown,
South Australia) during the Millennium Drought, which occurred between 2005 and 2010, adversely
contributed to more than 162 riverbank collapse-related incidents and a long-term metastable
condition to the riverbanks along the Lower River Murray, which have recently been considered as
the dominating factors inducing bank collapse. A number of studies have recently been undertaken
by the authors to examine the causes of these historical collapses, as well as developing a framework
to identify regions along the Lower River Murray that have a potentially higher risk of riverbank
collapse. This report summarises these riverbank collapse studies.

Back analyses of some of the larger past failures were undertaken at a number of sites, namely Long
Island Marina near Murray Bridge; East Front Road between Mannum and Younghusband;
Whitesands near Tailem Bend; and Woodlane Reserve. Based on government inventories, the
collapsed riverbank sections were identified and vectorised using visual interpretation within the
ArcGIS geographic information system (GIS) framework. With the aid of high-resolution aerial
photographs, digital elevation models (DEMs) and ArcGIS, the two-dimensional (2D) models of past
collapsed riverbank were established and analysed. A three-dimensional model (at Long Island
Marina) was also analysed. The integration of GIS with high-resolution spatial data facilitates the
process of the identification of collapsed regions, model geometry development and the calculation
of the dimensions of the collapsed regions.

From the GIS and DEM data, the topographies of the riverbanks were input into the SVSlope slope
stability analysis program, which was in turn linked to the SVFlux groundwater seepage modelling
software in order to perform transient back-analysis of riverbank stability. The operating shear
strengths of the soils were determined using back-analyses and these were found to be consistent
with those obtained from laboratory and field testing, and reported previously. The studies have also
demonstrated that the riverbank collapses can be accurately and reliably predicted by using the
proposed frameworks. Both the adopted process of 2D and 3D stability modelling yielded excellent
predictions of the collapses when compared against the recorded dates and dimensions of the failed
regions.

In summary, the studies have shown that the unprecedented low river levels are the main cause of
the past riverbank instability incidents. Transient analyses have shown that river fluctuations, rather
than climatic factors, dominate the likelihood of riverbank collapse along the Lower River Murray.
Sudden or rapid drawdown events (ratios of between –0.006 to –0.5 m per day) have been simulated
and it has been observed that such drawdown events may lead to riverbank instability. Furthermore,
extreme rainfall events (e.g. rainfall exceeding 120 mm per day, whilst unlikely but plausible)
coinciding with river levels at between –0.2 m to –0.5 m AHD, are also likely to trigger riverbank

Page 8 of 77
Riverbank Collapse Task 5 Report

collapses. Finally, this report also presents a framework for the preliminary identification of regions
along the Lower River Murray which potentially demonstrate high susceptibility to riverbank collapse.

Page 9 of 77
Riverbank Collapse Task 5 Report

1 Introduction
In general, bank retreat is the combined result of various influencing factors, such as bank geometry
(i.e. topography, slip surface and material boundary locations), riverbank materials, external loads
acting on the bank (i.e. vegetation and infrastructure) and pore water pressure (Filz et al., 1992).
Back-analysis is an analytical methodology to assess the status of inherent influencing factors on a
failed or failing slope (Filz et al., 1992). For slope stability analyses, at the time of failure, the details
of influencing factors and the ways they interacted are often ambiguous (Abramson et al., 2002).
Based on the assumption that the factor of safety (FoS) was equal to unity at the time of failure, a
back-analysis calculation can be undertaken and a series of relevant parameters obtained for a
specific location (Abramson et al., 2002).

Geographic information system (GIS) technology has greatly facilitated landslide research by providing
various functions for spatial data management. The GIS-based technique is comprised of data
capturing, handling, processing, analysing, integrating, simulating, visualising and modelling (Carrara
et al., 1991a, 1995b; Guzetti et al., 1999; Wang et al., 2005). GIS-based geotechnical models simplify
the process of quantifying the spatially distributed parameters influencing collapse (Xie et al., 2004).
By processing and integrating high-resolution, remotely-sensed data (i.e. satellite imagery, aerial
photos and digital elevation models [DEMs]), GIS-based geotechnical models have been used to
assess bank stability at relatively large scales (Hong et al., 2007). As a result, the GIS method has been
adopted throughout the study.

The stability of the riverbank is determined by means of the SVSlope (SoilVision, 2009b) slope stability
analysis program and expressed in terms of the Factor of Safety (FoS). The FoS is determined by
dividing the forces resisting movement by the forces driving movement. By this definition, a
geotechnical structure, or in the present study a riverbank, with a FoS of exactly 1 will support only
the design load and no more and will effectively be at the limit of stability. Any additional load, or
reduction in strength, will cause the riverbanks to collapse. Section 2 of this report outlines the
integration of GIS and SVSlope to analyse riverbank stability at Long Island Marina and Tailem Bend.

To understand the underlying physical mechanics of the riverbank collapses along the lower reaches
of the River Murray, as well as, to quantify the effects of river level fluctuation and climatic influences,
such as rainfall, wind, mean temperature and evaporation on riverbank stability, back-analysis was
undertaken to model 4 historical and representative riverbank collapses. Four very high-risk sites
identified by Sinclair Knight Merz (SKM) (2010a) namely, East Front Road near Mannum; Woodlane
Reserve; River Front Road near Murray Bridge; and White Sands, were selected for these analyses.
The back-analyses were undertaken by combining the GIS method, slope stability analysis using limit
equilibrium (SVSlope) and finite element simulation of groundwater seepage flow (SVFlux [SoilVision,
2009a]). The methodologies and the results of the transient analyses are presented in detail in Section

Page 10 of 77
Riverbank Collapse Task 5 Report

3 of this report. The results are shown as transient FoSs, that is fluctuation of FoS with historical river
level variations and climatic data.

Section 0 presents the modelling of the riverbank collapse incident, in both 2D and 3D, which
occurred at Long Island Marina, Murray Bridge, South Australia, on 4 Feb. 2009. The influence and
sensitivity of river level fluctuations and climatic factors on riverbank stability are examined in detail.
From the results of these analyses the dominant triggers of riverbank collapse were identified. Both
2D and 3D stability modelling yielded excellent predictions of the collapse when compared against the
recorded date and dimensions of the collapsed region. The operating geotechnical parameters were
obtained and are consistent with those derived from field investigations reported by Jaksa et al.
(2015). It showed river fluctuations, rather than climatic factors, are the triggers of riverbank
collapses along the Lower River Murray. Events such sudden or rapid drawdown and extreme rainfall
were also simulated.

Further to the Task 4 report by Hubble and De Carli (2015), Section 0 identifies locations of deep holes
in the channels of the Lower River Murray from Blanchetown to Wellington by analysing DEMs using
the geoprocessing framework in ArcMap. Hubble and De Carli (2015) suggested that deep holes are
formed due to either: (a) bedrock constriction and pronounced narrowing of the channel cross-
section; or (b) large outcrops of bedrock which protrude up from the floor of the channel. Such
geological features have generated erosive flow patterns during periods of higher flow that have
eroded or scoured deep holes in the channel and over-steepened the channel margins and banks and
past large failures were often associated with the these features. Deep holes are more common
downstream of Mannum, where most of the large failures have occurred. Riverbanks near Thiele
Reserve have been extensively analysed and the results showed that riverbanks with slope angles
steeper than 40° are vulnerable and could collapse in the longer term, but the failure types are most
likely to be shallow or planar, implying that, if they occur, relatively small volumes of soil will collapse
into the river. Finally, the study showed that the slope inclination map is a very useful indicator for
identifying riverbank instability near Thiele Reserve and will thus be explored further in future work.

Page 11 of 77
Riverbank Collapse Task 5 Report

2 GIS-based Analysis of Riverbank Collapses at Long Island


Marina and Tailem Bend

2.1 Introduction
This section demonstrates the use of back-analysis to assess the susceptibility of riverbank collapses
of several failed sites at different river levels at Long Island Marina, Murray Bridge, South Australia. It
presents the first effort in this research program to back-analyse the Long Island Marina failure, and
was published previously by Liang et al. (2012). Section 3 builds upon these analyses and examines
other sites where riverbank failures have occurred between 2005 and 2010 and Section 4 further
extends and refines the Long Island Marina analyses and explores various river level and climatic
scenarios.

The study area examined in the present section is located along the Lower River Murray near the Long
Island Marina, Murray Bridge, South Australia, as shown in Figure 1. The historical inventory of
riverbank collapse incidents (WaterConnect 2015), which is maintained by the South Australian
Department of Environment, Water and Natural Resources (DEWNR) (formerly DFW), documents 9
riverbank collapse-related incidents occurred between 2008 and 2011. Of these, 4 were relatively
minor (i.e. bank cracking, tree leaning and collapse), however, the remaining 5 were significant and
are located adjacent to Long Island Marina, as shown in Figure 2(a).

Figure 1: Location of the study area.

For the purpose of back analysis, the collapsed sections of riverbank need to be identified with
relatively high accuracy. With reference to a series of imprecise geographical coordinates of the
collapsed bank sections from the DEWNR inventory, under the ArcGIS framework, a visual
interpretation method was adapted using high-resolution aerial images taken at different periods to
identify the collapse region. Based on the visual comparison of the aerial images taken in 2008, with
a 5 m resolution, and in 2010, with a 0.5m resolution, the collapsed region was vectorised, as shown
in Figure 3 by the dotted hatching.

Page 12 of 77
Riverbank Collapse Task 5 Report

There are shortcomings, however, when attempting to identify collapsed regions by means of visual
interpretation. Specifically, it is extremely difficult to verify by eye whether a section of riverbank
collapsed or is submerged beneath a higher river level. In order to resolve this issue, a digital
elevation model (DEM) acquired in 2008 was used to validate the elevation of the section prior to
collapse. By comparing the elevations from the DEM with the river levels at the date that the aerial
photograph was taken (river level data were obtained from the observation point at Murray Bridge),
collapsed regions were more accurately identified. Furthermore, the loss of large riparian plants and
the construction of new infrastructure (e.g. jetty), observed from later aerial images, are also helpful
in identifying the collapsed regions.

Figure 2: Long Island Marina study site: (a) locations of 5 significant failures; (b) location plan of in situ testing
and recorded collapses; and (c) distribution of bank cross-sections in 3D perspective.

Figure 3: Examples of visual interpretation of 2008 and 2010 aerial images using ArcGIS at:
(a) Murray Bridge; and (b) Tailem Bend.

Page 13 of 77
Riverbank Collapse Task 5 Report

2.2 Background and Methodology


Riverbanks similar in material, groundwater and stratigraphy may behave significantly differently
depending on their particular topographies (Abramson et al., 2002). Throughout this study, the
geometries of the riverbanks were extracted from two DEMs (0.5 m and 5 m resolution, between the
years 2007 and 2010), which were provided by DEWNR (WaterConnect, 2015), within the ArcGIS
framework. As shown in Figure 2(b), 26 2D riverbank cross-sectional models were established and
represented by 26 polyline slices. Each polyline consisted of 401 points in a line with a 10 m interval
between neighbouring points. In order to conduct the back analyses, 5 cross-sectional models were
selected to coincide with collapsed regions, which were vectorised previously. Based on the bilinear
interpretation method, the elevation values were extracted from the DEMs and assigned to 401
points with singular points avoided.

Stratigraphic and geotechnical data for this research were collected from a series of site investigations
(boreholes, and cone penetration, dilatometer, vane shear, standard penetration, pocket
penetrometer and laboratory tests), performed by Sinclair Knight Merz (SKM, 2010a). Generally
speaking, the soil profile comprises a 0.8 – 1.2 m thickness of silty sand (SM/SC) overlaying a silty clay,
(unit weight of 18 ± 1kN/m3, internal angle of friction of 28 ± 2°, cohesion of 2 ± 2 kPa); 11 – 20 m
thickness of dark grey, very soft, silty clay (CH) (unit weight of 16 ± 1 kN/m3, cohesion of 10 ± 5 kPa
[10 kPa at the top, maximum of 25 kPa increasing by 1.25 kPa/m]); and a medium dense clayey sand
(Monoman Sand) (SC/CL) layer (unit weight of 17 ± 1 kN/m3, internal angle of friction of 30 ± 2° and 2
± 2 kPa cohesion).

Pore water pressure affects the riverbank stability by altering the shear strength of bank material
(reduces the effective stress) and the self-weight of the soil mass (Sharma et al., 2002). It is well
understood that positive pore water pressure, which plays a significant role in drawdown failures,
reduces the effective shear strength of soils (Budhu and Gobin, 1995). In contrast, negative pore
pressure, or soil suction, offers apparent cohesion, which stabilises the bank and manifests steeper
banks. Casagli et al. (1997) stated that when the bank materials become saturated, collapses are
more likely to occur. However, due to seasonal variations, negative pore water pressure do not
contribute to long term stability (Sterrett and Edil 1982).

Fluctuations in river level directly affect the flow of water in and out of the bank and the pore water
pressures within the bank (Green, 1999). Generally, over the last 2 decades or so, the river levels at
Murray Bridge have fluctuated between 1.5 – 1.7 m AHD (Australian Height Datum), as shown in
Figure 4. In general, the river level has remained relatively constant until the end of 2006, where it
fell dramatically until October 2009, after which it has returned to its pre-2006 levels.

Page 14 of 77
Riverbank Collapse Task 5 Report

1.5

1
Water Level (m AHD)

0.5

-0.5

-1

-1.5

-2
01-Dec-86

01-Dec-87

01-Dec-88

01-Dec-89

01-Dec-90

01-Dec-91

01-Dec-92

01-Dec-93

01-Dec-94

01-Dec-95

01-Dec-96

01-Dec-97

01-Dec-98

01-Dec-99

01-Dec-00

01-Dec-01

01-Dec-02

01-Dec-03

01-Dec-04

01-Dec-05

01-Dec-06

01-Dec-07

01-Dec-08

01-Dec-09

01-Dec-10
Date

Figure 4: River Murray water level at Murray Bridge 1/12/1986 to 11/07/2011 (based on DFW, 2010).

Applied external loads are also known to affect slope stability. The external loads acting on
riverbanks consist of two main categories: riparian plants and manmade structures. Considering the
potential slip modes and the horizontal and lateral extent of the collapse, the impacts of failure were
mainly determined by the dead weights and the distributions of the external loads (distances to the
bank line).

The effects of riparian plants on riverbank stability are well known (e.g. Hickin, 1984, Thorne, 1990).
The vegetation alters bank hydrology, flow hydraulics and the bank’s geotechnical properties
(Abernethy and Rutherfurd, 2000). However, the effects of plants change with the seasons and their
life cycle and, hence, are difficult to predict and integrate into bank stability analysis (Thorne and
Osman, 1988; Abernethy and Rutherfurd, 2000). The self-weight of large vegetation, such as trees,
and the effects of wind on the trees’ canopies, add load to the riverbank, which in turn, destabilises
the bank. Combining the beneficial hydrological and mechanical stabilising effects from root suction
and root matrix reinforcement, with the detrimental, additional load provided to the bank results in
vegetation providing only a marginal influence on bank stability (Abernethy and Rutherfurd, 2000).

The location and the self-weight of manmade structures affect riverbank stability by imposing a
surcharge on the bank. In the study area, a site inspection indicates that approximately 33% of
houses are located quite close (less than 10 m) to the riverbank line. The load exerted by a house to
the bank is roughly 20 kPa, which consists of the weight of the roof, walls and floor, as well as the live
loads, as shown in Figure 5.

Using the data discussed above, a series of back-analyses were conducted using the limit equilibrium
method within SVSlope (SoilVision, 2009b) assuming that a FoS of unity represents incipient failure.

Page 15 of 77
Riverbank Collapse Task 5 Report

Due to the distances between the site investigations and the actual locations of the collapse sites, the
soil parameters collected from borehole logs were only applied as initial values in the FoS calculation.

Figure 5: Schematic of the locations and the self-weight of the structures on the riverbank.

The results of site investigations suggest that deep-seated rotational slips due to bank toe erosion in
low flow situations, as well as slab failures, are the two dominant riverbank retreat modes along the
Lower River Murray (Tajeddin et al., 2010; Hubble and De Carli, 2015). The deep-seated rotational
failures typically occurred in the silty clay layer (between –5 to –20 m AHD), an example of which is
shown in Figure 6.

Figure 6: Minimum FoS and potential slip surface of a deep-seated rotational failure obtained from SVSlope at
Location No. 21 when the water level was at 0 m AHD.

Page 16 of 77
Riverbank Collapse Task 5 Report

The results of the back analyses are summarised in Figure 7: two geotechnical models were adopted
for the silty clay layer – Model 1: Unit weight = 16 kN/m3; ctop = 6.4 kPa; cmax = 25 kPa; cratio =
1.25kPa/m; Model 2: Unit weight = 16 kN/m3; ctop = 10 kPa; cmax = 25 kPa; cratio = 0.87 kPa/m. The
geotechnical model associated with the results of the site investigation is indicated in Figure 7 by “Site
Investigation Model.”

Figure 7: Back analyses using three geotechnical models.

As mentioned previously, there is widespread agreement that river drawdown adversely affects
riverbank stability (e.g. Springer, 1981; Mayo, 1982; Thorne, 1982; Springer et al., 1985; Dahm et al.,
1988; Arnott, 1994; Budhu and Gobin, 1995). In general, there are two types of drawdown: short-
and long-term, both of which remove the lateral support from the channel and destabilise the bank.
The difference being, in short-term drawdown, the pore water pressures within the bank do not have
time to equilibrate with the drawn-down river level (Budhu and Gobin, 1995). The circumstances at
Long Island Marina suggest that long-term drawdown is more relevant, which leads to permanent
changes in the properties and behaviour of bank assets and soils (SKM 2010).

With reference to the river level records at Murray Bridge during 2011, the river fluctuated between
0.5 and 1.2 m AHD, and rarely drew down below 0.5 m AHD. Based on this, a reasonable assumption
for a low river level threshold, in foreseeable future, is expected to lie between 0 and 0.5 m AHD,
although river management authorities are likely to seek to maintain the river at a higher level than
this. In order to examine the susceptibility of the riverbank to collapse, simulations were performed
at two river levels (0 and 0.5 m AHD) with two geotechnical models (Figure 8).

With reference to the River Murray, a long term FoS of 1.5 has been adopted as the minimum
acceptable FoS, accounting for human safety and the protection of assets (DFW, 2010). As a
consequence, FoS ≤ 1 is adopted as “Unstable”, which indicates additional stabilisation is needed for
stability; 1 < FoS ≤ 1.25 is assumed to be “Quasi stable”, indicating that minor destabilising factors can

Page 17 of 77
Riverbank Collapse Task 5 Report

cause instability, such as additional surface loading, severe boat wash, river drawdown and high winds
affecting trees; 1.25 < FOS ≤ 1.5 is adopted as “Moderately stable”, implying that moderate
destabilising factors can cause instability, such as construction and extensive river drawdown; and
FOS > 1.5 was adopted as “Stable”, denoting that only major destabilising factors are likely to cause
instability.

2.3 Results
The results of the back-analysis at 0 m AHD river level are shown in Figure 8(a). It can be seen that
Cross-sections 5, 13 and 21 are about to collapse; 6, 7, 19 and 22 are at the limit of failure; 9, 10, 11,
12, 15, 18, 20 and 23 are moderately stable; and 8, 14, 16, 17, 24 and 25 are stable.

When the river levels are maintained at 0.5 m AHD, as indicated in Figure 8(b), it is observed that
Cross-section 13 is about to collapse; Cross-sections 5, 6, 7, 19, 21 and 22 are at the limit of failure; 9,
10, 11, 12, 20 and 23 are moderately stable; and 8, 14, 15, 16, 17, 18, 24 and 25 are stable.

(a)

(b)

Figure 8: Factors of Safety of neighbouring cross sections (0 and 0.5 m AHD).

Page 18 of 77
Riverbank Collapse Task 5 Report

As shown in Figure 9, the risk of riverbank instability is, in general, reduced by increasing river levels.
Hence, when the river level rises to 0.5 m AHD, a relatively large proportion of the banks examined
present as quasi stable rather than moderately stable.

Figure 9: Predictions of riverbank susceptibility with river levels at (a) 0 m AHD and (b) 0.5 m AHD.

The low river levels in the Lower River Murray during the Millennium Drought have caused the banks
to remain in a long-term, metastable condition since 2007. The study site was adjacent to Long Island
Marina at Murray Bridge in South Australia where a series of riverbank collapses have occurred
relatively recently. With high-resolution LIDAR (light detecting and ranging) data and aerial photos,
this study has vectorised the collapsed riverbank sections by visual interpretation and 2D back-
analyses have been performed on 5 cross-sections in the collapsed region using GIS, site investigation
data and the slope stability program, SVSlope. In addition, a series of sensitivity analyses has been
conducted at two different water levels: 0 and 0.5 m AHD at 21 cross-sections using the results of
back-analyses.

2.4 Summary
The results of this work have shown that, when the river level is low, a section of the riverbank is
close to collapse, whereas a number of cross-sections adjacent to the study site are susceptible to
riverbank collapse and require further investigation. In addition, it has been observed that increased
river levels generally stabilise the riverbanks but to a limited extent. Several remedial works may
need to be conducted when the river level is predicted to fall for an extended period. Later sections
will incorporate climatic factors into the riverbank stability assessment at this and several other sites
along the Lower River Murray.

Page 19 of 77
Riverbank Collapse Task 5 Report

3 Back Analysis of Historical Riverbank Collapses

3.1 Introduction
For the purpose of understanding the underlying mechanics of the riverbank collapses along the
lower reaches of the River Murray, as well as quantifying the effects of river level fluctuation, rainfall
and evaporation on riverbank stability, back-analyses have been undertaken to model 4 historical and
representative riverbank collapses. These riverbank collapses took place at 4 high-risk sites identified
by Sinclair Knight Merz (SKM, 2010a). These sites, which include Mannum (MN); Woodlane Reserve
(WR); River Front Road, Murray Bridge (RFR); and White Sands (WS), are associated with 50 of the
total 162 reported riverbank collapse-related incidents, according to the inventory of the South
Australian Department of Environment, Water and Natural Resources (DEWNR) (formerly DFW)
between the years 2007 and 2010 (WaterConnect, 2015). The work reported in this section was
recently published by Liang et al. (2015a). The geographic locations and nature of these historical
riverbank collapse-related incidents associated with these four sites are summarised in Table 1 and
Figure 10.

Table 1: Nature of the historical riverbank collapse-related incidents at 4 study sites.

Sites Bank collapse Bank cracking Tree leaning and collapse Levee problem
MN 6 7 7 0
WR 3 3 0 0
RFR 8 3 6 2
WS 3 0 3 1

At MN, a section of the riverbank slumped into the river and some areas at MN were observed to
exhibit cracking along the riverbank, suggesting impending failure zones had developed. At WR, a 45
× 14 m section of riverbank collapsed, destroying a pumping station. At RFR, which was discussed in
the previous section, an unprecedented period of dry conditions and low flows induced a significant
section of riverbank (60 × 20 m, 70,000 m3) to collapse into the river, taking with it three unoccupied
vehicles and several trees. At WS, two large riverbank collapses were reported: 20 × 6 m on 14
February 2009 and 25 × 4 m on 22 April 2009 (SKM 2010a).

Stability failures are often back-analysed, as mentioned earlier, to estimate the in situ shear strength
at the time of the collapse (Abramson et al. 2002). An alternative approach to back-analysis is the
determination of soil shear strength by means of undertaking laboratory and/or in situ testing.
However, both laboratory and in situ testing present limitations. Laboratory testing, for example, is
associated with a number of shortcomings, such as the sample disturbance, in which the structure of
the soil has been substantially disturbed due to sampling, such that subsequent laboratory tests are
not representative of in situ conditions. Moreover, to estimate accurately the shear strength of soil
materials, the field conditions need to be accurately replicated in the laboratory, including the

Page 20 of 77
Riverbank Collapse Task 5 Report

following states prior to failure: effective normal stress acting on the failure surface; pre-existing
shear deformation; and the drainage conditions during shear (Tang et al. 1999).

(MN)

(WR)

(RFR)

(WS)

Figure 10: The geographical locations of the four studied sites and major riverbank collapses in the past, as
well as the proximities of the borehole and river level observation stations.

Page 21 of 77
Riverbank Collapse Task 5 Report

On the other hand, the examples of the limitations of in situ testing include the uncertainty of the
models used to translate the field measurements into their associated soil properties, as well as the
human factors in operating the testing equipment and conforming to the standardised test
procedures. For these reasons, back-calculated soil shear strengths have advantages over laboratory-
determined data, because they represent a large scale of soil mass over slip surfaces and they were
determined from an in situ state (Gilbert et al., 1998). A back-analysis assumes: (i) a slope geometry
prior to failure; and (ii) a factor of safety (FoS) is equal to unity that is the condition of the failure to
have occurred according to the limit equilibrium method to estimate the in situ shear strength at the
time of failure (Okui et al., 1997; Gilbert et al., 1998; Urgeles et al., 2006; Bozzano et al., 2012; Harris,
Orense & Itoh, 2012; Wang et al. 2013).

3.2 Methodology
In this section, an efficient framework for the back-analysis of riverbank collapses is adopted at the 4
aforementioned sites. The framework employs a finite element analysis-based transient water model
to evaluate the dynamic distribution of pore water pressure under different circumstances of rainfall,
evaporation and river level fluctuations. Subsequently, limit equilibrium slope stability analyses are
performed to back-analyse the failure. The work presented in this section aims to: (i) examine the 4
aforementioned riverbank collapses using 2D cross-sectional models; and (ii) determine the soil shear
strength profiles from the back-analyses.

In order to determine the location of the collapses with a relatively high degree of accuracy, visual
interpretation of high-resolution aerial images has been implemented in conjunction with DEWNR’s
inventory of historical riverbank failures (shown as green squares in Figure 11). For each riverbank
collapse, two aerial photographs have been used to identify the extent of the collapse regions. Figure
11(a), (c), (e) and (g) were acquired in March 2008 with a 0.5 metre resolution; whereas Figure 11(b),
(d), (f) and (h) incorporate a 0.2 metre resolution which were acquired after the recorded collapses
(May 2010). As shown in Figure 11, the examined regions of collapse have been vectorised with
dotted areas within ArcGIS, and have been linked to the closest DEWNR collapse record (incident IDs
23 (MN), 47 (RFR), 108 (WS), 113 to 115 (WR) in the Riverbank Collapse Hazard Incident Register
[WaterConnect, 2015]).

A transient, unsaturated flow-based riverbank stability model has been developed in two dimensions
to facilitate back-analyses at the 4 aforementioned sites along the lower reaches of the river. The
model incorporates riverbank geometry, geotechnical properties, river level variation, and rainfall and
evaporation. Back-analyses have been performed to obtain the closest match between the predicted
and actual date of failure, while comparing the predicted failure geometry with the high-resolution
aerial images.

Page 22 of 77
Riverbank Collapse Task 5 Report

(a) (b)

(c) (d)

(e) (f)

(g) (h)

Figure 11: Adopted visual interpretation method of high-resolution aerial images: (a), (c), (e) and (g) are aerial
photographs acquired in March 2008 at MN, WR, RFR and WS, respectively; (b), (d), (f), and (h) are aerial
photographs acquired in May 2010 at MN, WR, RFR and WS, respectively.

Page 23 of 77
Riverbank Collapse Task 5 Report

An elevation comparison method has been employed at each of the sites to assist with visual
interpretation. Specifically, LIDAR (light detection and ranging) DEMs obtained in 2008 [Figure 12(a)]
were used to verify the elevation of the examined collapses with the LIDAR DEMs obtained in 2010
[Figure 12(b)], which include the elevation of the river level. For example, point A in Figure 12(a),
where the elevation of the bank was at 1.484 m AHD prior to the collapse, is compared with point B,
where the river level was at –0.45 m AHD subsequent to the failure. By comparing the DEMs, each of
the examined regions [Figure 11(b), (d), (f) and (h)] is confirmed as collapsed rather than simply
submerged beneath a higher river level.

(a) (b)
Figure 12: Example of adopted elevation comparison method on DEMs at WR (a) 1 m resolution DEM acquired
in 2008; (b) 0.2 m resolution DEM acquired in 2010).

With reference to the DEWNR riverbank collapse inventory (WaterConnect, 2015), visual
interpretation was implemented on high-resolution aerial images to examine the region of collapse
within ArcGIS. The 2D cross-sections that coincided with the collapse regions have been used to
analyse the evolution of the FoS of the riverbank at the 4 sites. The groundwater flows in the
saturated and unsaturated zones within the riverbank have been simulated using the finite element
method within the PC-based program SVFlux (SoilVision 2009a). The 2D limit equilibrium slope
stability calculations have again been performed using SVSlope (SoilVision 2009b). Specifically, in
SVFlux, the variations in rainfall, evaporation and river level fluctuations are entered to model the
dynamic distribution of pore water pressure within the riverbank. These are subsequently imported
into SVSlope to assess the stability of the riverbank against time. By considering the vertical extents
of the collapses, back-analyses have been performed on the layer of high plasticity Silty Clay (CH), as
outlined below, using the Unsaturated Fredlund soil model (Fredlund et al. 1996) informed by
geotechnical investigations performed near each site (Jaksa et al., 2015).

In this section, the geometries of the riverbank associated with the 4 sites have been obtained from
DEMs using ArcGIS. As shown in Figure 12, each of the cross-sections initially comprised of 401 points,

Page 24 of 77
Riverbank Collapse Task 5 Report

with a 10 m lateral interval between each neighbouring point. Again, based on the bilinear
interpolation method, the elevation data have then extracted from the two DEMs (a 1 m resolution
DEM acquired in 2008 and a 0.2 m resolution DEM acquired in 2010, as mentioned above) and
assigned to those 401 points for each cross-section, with singular points avoided. As indicated in
Figure 11, the locations of the 4 cross-sections coincide with and cross each examined region of
collapse. This process generated high-resolution cross-sections, which provide effective modelling of
the riverbanks in both SVFlux and SVSlope.

In October and November 2009, geotechnical investigations were performed by SKM at three sites,
including MN (3 boreholes and 1 piezocone [CPTu]), WR (2 boreholes and 5 CPTus) and RFR (2
boreholes and 13 CPTus) (SKM 2010b). Based on the data from investigations, summarised in Table 2
and Figure 4, the geotechnical models have been developed at each of the 4 sites. The WS and RFR
sites share a common geotechnical model, which was obtained from the RFR geotechnical
investigation, whereas the models for MN and WR have been established from their respective
investigations, as indicated above.

For the purpose of modelling the variation of soil suction under saturated and unsaturated conditions,
soil water characteristic curves (SWCCs) are used in each of the soil layers within SVFlux. The SWCC
parameters are estimated from the particle size distributions (PSDs) in each soil layer using the Zapata
method (Torres Hernandez, 2011), which incorporates the weighted Plasticity Index (wPI). Adopting
Equation (1), wPI is determined based on the PSDs from the borehole soil samples (Table 2).

wPI = IP × P200 ⁄100 (1)

where IP is the plasticity index and P200 is the percentage of soil passing size #200 (US sieve size #200,
which is equivalent to a 0.074 mm aperture).

The saturated hydraulic conductivity, Ksat, is derived from a CPTu pore pressure dissipation test
performed at 5 m depth in Murray Bridge, as shown in Figure 13(b). Based on relationships between
heavily consolidated silts and clays, monotonically decreasing excess pore pressure with time
response and the saturated hydraulic conductivity (Burns and Mayne 1998) for the CH and CL layers,
Ksat is found to equal approximately 9.9 × 10–5 m/day, as shown in Table 2 and Figure 13(b). In
unsaturated soil, the unsaturated hydraulic conductivity, K, varies with respect to matric suction, and
is calculated indirectly from the Fredlund and Xing (1994) estimation associated with Ksat.

The river level data used in this study are obtained from 4 observation stations: A4261161, A4260547
(WaterConnect 2014), A4261162 and A4260522 (MDBA 2014), as shown previously in Figure 10. The
climatic data, which include mean daily rainfall, mean daily temperature, evaporation and humidity,

Page 25 of 77
Riverbank Collapse Task 5 Report

at the 4 sites are collected from the Bureau of Meteorology (2014). In order to back-analyse the
known riverbank collapses under the combined influence of river level fluctuation, rainfall and
evaporation, a one-month historical record was adopted prior to the recorded date of each riverbank
collapse.

As shown in Figure 14, unexpected low inflows, combined with evaporation of the lower lakes,
resulted in the daily river levels remaining between –1.1 to –0.8 m AHD at these 4 aforementioned
sites. More specifically, compared with the relatively constant daily river levels at WR and RFR,
significant high flow events were observed at MN and WS in the last 7 days of the month.
Approximately 30 mm and 35 mm rainfall were observed at MN and WS, respectively, however, the
rainfall at WR and RFR sites was negligible.

Table 2: Soil properties for saturated and unsaturated flow modelling.

Elevation Ksat θsat ρ w IP γ P200


Layer
(m AHD) (m/day) (%) (t/m3) (%) (%) (kN/m3) (%)
Mannum (MN)
Silty/Clayey Sand
3.5 to 1 13.51 45.2 1.7 17.1 2 20 ± 1 21
(SM/SC)

Silty Clay (CH) 1 to –1 9.9 x 10–5 59.8 1.4 26.6 16 17 ± 1 20

Silty Clay (CH) –1 to –5 9.9 x 10–5 80.7 0.95 79 42 17 ± 1 15

Clayey Sand (SC) > –5 0.187 49.7 1.6 59 38 20 ± 1 20


Woodlane Reserve (WR)
Sand (SP) 2.5 to 2 15.21 62 1.35 35.4 40 17 ± 1 78

Silty Clay (CH) 2 to –1 9.9 x 10–5 76 1 56 41 17 ± 1 37

Clayey/Silty Sand
–1 to –5 0.187 57 1.7 48.6 17 20 ± 1 29
(SC/SM)

Sandy/Silty Clay (CL) –5 to –11 9.9 x 10–5 52 1.6 22 14 20 ± 1 68

Silty Sand/ Gravel


> –11 0.187 51 1.6 22 3 20 ± 1 18
(SM)
River Front Road (RFR) and White Sands (WS)
Silty/Clayey Sand
1 to 0 13.51 52.7 1.25 42 30 18 ± 1 35
(SM/SC)

Silty Clay (CH) 0 to –20 9.9 x 10–5 63.2 1.01 61 50 16 ± 1 67

Clayey Sand/Sandy
> –20 0.187 54.6 1.23 44 50 17 ± 1 59
Clay (SC/CL)

Page 26 of 77
Riverbank Collapse Task 5 Report

(a)

(b)
Figure 13: Results of CPTu sounding (a) profile; and (b) pore pressure dissipation test results.

Page 27 of 77
Riverbank Collapse Task 5 Report

(a) (b)

(c) (d)
Figure 14: Daily river levels and daily rainfall recorded at (a) Mannum (MN) site in April 2009;
(b) Woodlane Reserve (WR) site in February 2009; (c) River Front Road, Murray Bridge (RFR)
site in February 2009; and (d) White Sands (WS) site in April 2009.

The dates of each riverbank collapse incident have been obtained from the most relevant DEWNR
record, and the back-analyses are benchmarked against these dates. The riverbank profiles at the 4
sites were modelled previously by SKM (2010a) and Coffey (2012) using separate soil layers informed
by borehole logs from SKM (2010a, 2010b, 2010c, 2010d, 2010e). Specifically, the static SKM and
Coffey analyses, as summarised in Table 3 and Figure 15 to Figure 18, incorporated effective stress
parameters for the Fill and SC layers using the Mohr-Coulomb failure criterion, while total stress
parameters were adopted in the CH layers using the following depth-dependent, undrained model:
linearly-increasing cohesion with depth, with ctop quantifying the cohesion (kPa) at the upper layer
boundary, cratio representing the gradient of increasing cohesion with depth and capped at a
maximum value of cmax. Each of the values recommended by SKM (2010a) and Coffey (2012) were
inferred from laboratory and in situ test results obtained from their respective geotechnical
investigations. However, in order to accommodate the effects of positive and negative pore water
pressures in the partially-saturated riverbanks, an unsaturated, effective stress analysis, based on the
Unsaturated Fredlund model (Fredlund et al. 1996) is performed on the Fill and CH layers in the
present paper, as shown by Equation (2).

𝜏 = 𝑐 ′ + (𝜎𝑛 − 𝑢𝑎 )𝑡𝑎𝑛∅′ + (𝑢𝑎 − 𝑢𝑤 )[𝜃(𝑢𝑎 − 𝑢𝑤 )⁄𝜃𝑠 ]𝑏 𝑡𝑎𝑛∅ (2)

Page 28 of 77
Riverbank Collapse Task 5 Report

where 𝜃(𝑢𝑎 − 𝑢𝑤 ) = the volumetric water content at any suction; 𝜃𝑠 = saturated volumetric water
content; b = a fitting parameter that has a value close to unity for sands and increases with plasticity.
At each site, the geotechnical properties of the high plasticity, clay (CH) layers are varied marginally,
with respect to the results of the corresponding site investigations, until the FoSs of unity were
obtained in close proximity to the recorded dates of collapse. As shown in Figure 15 to Figure 18,
values of FoS = 1.0 were obtained from the most appropriate Unsaturated Fredlund soil model as
summarised in Table 3. For example, at the River Front Road (RFR) site at Murray Bridge, the soil
model for the CH layer adopted c' = 0 kPa; φ' = 22°, which aligns reasonably well with the single CU
triaxial test performed by SKM (2010b) (φ' = 27°).

Table 3: Geotechnical models of the clay layer obtained from back analyses.

Effective stress
Undrained parameters for CH Undrained parameters for CH
Sites parameters from back
layer from SKM (2012) layer from Coffey (2012)
analysis
cu-B = 50 ± 10 (kPa)
MN cu-B1 = 17.5 ± 2.5 (kPa) c' =0, '=23°
cu-B2 = 14± 2 (kPa) cu-top = 5.5 (kPa)

WR cu = 20 ± 5 (kPa) cu-ratio = 1.25 (kPa/m) c' =0, '=20°

cu-top = 10 ± 5 (kPa) cu-max = 25 ± 5 (kPa)


RFR cu-ratio = 1.25 (kPa/m) c' =0, '=19°
cu-max = 25 ± 5 (kPa)

WS N/A N/A c' =0, '=24°

3.3 Results
As mentioned above, the models are validated in two ways. Firstly, the predicted dates of bank
collapse are compared with the historical collapse dates from DEWNR inventory. As can be seen from
Table 4 and Figure 19 to Figure 22 the predicted dates compare very favourably with the historical
dates. This is not particularly unexpected, given that the values of φ' were varied until the predicted
and historical collapse dates were in good agreement. However, Figure 19 to Figure 22 are
particularly encouraging in that the FoS time series, prior to collapse, consistently plot above unity,
which is consistent with increasing pore water pressure prior to failure. In addition, the second
independent validation measure is the volume of the collapsed region. By scrutinising the high-
resolution aerial images (Figure 11), the dimensions of the collapsed regions can be determined, as
explained earlier. Table 4 also shows the predicted widths of the collapsed regions from the back-
analyses of the 4 sites. As can be seen, the predicted widths compare extremely well with those of
the actual collapses.

Page 29 of 77
Riverbank Collapse Task 5 Report

Figure 15: Riverbank stability analysis of the Mannum (MN) site on 21 April 2009.

Figure 16: Riverbank stability analysis of the Woodlane Reserve (WR) site on 26 February 2009.

Figure 17: Riverbank stability analysis of the River Front Road, Murray Bridge (RFR) site on 6 February 2009.

Page 30 of 77
Riverbank Collapse Task 5 Report

Figure 18: Riverbank stability analysis of the White Sands (WS) site on 23 April 2009.

Table 4: Model validation.

Width of collapsed
Recorded date of Predicted date of Predicted width of
Sites region from aerial
collapse collapse collapsed region
photos
MN Approx. Day 20 Day 23 8m 8m

WR Approx. Day 28 Day 26 18 m 17.7 m

RFR Approx. Day 4 Day 6 20 m 19.5 m

WS Approx. Day 22 Day 23 6m 5m

Figure 19: Riverbank collapse factor of safety time series for Mannum (MN) in April 2009.

Page 31 of 77
Riverbank Collapse Task 5 Report

Figure 20: Riverbank collapse factor of safety time series for Woodlane Reserve (WR) in February 2009.

Figure 21: Riverbank collapse factor of safety time series for River Front Road (RFR) in February 2009.

Figure 22: Riverbank collapse factor of safety time series for White Sands (WS) in April 2009.

Page 32 of 77
Riverbank Collapse Task 5 Report

3.4 Summary
In summary, back-analysis of riverbank collapse has been undertaken and applied to 4 high-risk sites
along the lower reaches of the River Murray. A GIS framework, incorporating LIDAR digital elevation
models and high-resolution aerial images, has been used to determine the riverbank geometries prior
collapse and cross-examine the actual collapsed regions from the inventory. Geotechnical data,
which have been used in the back-analyses, were obtained from the site investigations performed at
the 4 sites. A transient, unsaturated flow-based riverbank stability model was implemented using
SVSlope in conjunction with SVFlux.

The study has shown that:


(a) The results of back-analysed soil shear strengths at the 4 sites have shown very good
consistency with those proposed by the geotechnical consultant (SKM) commissioned to
undertake site investigations adjacent to the collapse sites;
(b) Model validation demonstrates the adopted framework provides reliable riverbank stability
predictions; and
(c) The integration of GIS with high-resolution spatial data facilitates the process of collapsed-
region identification, model geometry development and the calculation of the dimensions of
the collapsed regions.

Page 33 of 77
Riverbank Collapse Task 5 Report

4 Effects of River Level Fluctuation and Climate on Riverbank


Stability
This section presents the results of numerical modelling performed to examine the influence of river
level fluctuations and climatic factors on the occurrence of riverbank collapses. This work was
originally published by Liang et al. (2015c) but has been updated to include recent developments.

4.1 Background
The effects of climate and river level fluctuation on the stability of riverbanks have been extensively
explored and discussed by several researchers (Hooke, 1979; Springer, 1981; Thorne, 1982; Osman
and Thorne, 1988; Thorne and Osman, 1988; Casagli et al., 1999; Green, 1999; Pauls et al., 1999;
Rinaldi and Casagli, 1999; Lane and Griffith, 2000; Zhang et al., 2005; Jia, 2009; Li et al., 2011).
Compared with the assessment of landslides in mountainous regions, changes in pore water pressure,
in particular soil suction (negative pore water pressure), plays a more fundamental role in the stability
of riverbanks (Hooke, 1979; Thorne, 1982; Casagli et al., 1999; Rinaldi and Casagli, 1999; Abramson et
al., 2002). The pore water pressure significantly affects riverbank stability by changing the soil shear
strength (Section 2). As stated by Fredlund (2006), soil suction can vary from zero to approximately
1,000 MPa in the unsaturated (or vadose) zone (Bouwer, 1978) and is highly dependent on the
properties of the soil, hydrological conditions and soil–atmospheric interactions (Lane and Griffith,
2000).

Rainfall has long been recognised as one of the most significant factors responsible for initiating slope
failures in many tropical or subtropical regions (Zhang et al., 2011). Generally, rainfall-induced slope
failures are observed as shallow failures; however deep-seated rotational failures are also reported
(Zhang et al., 2011). According to Rahardjo et al. (2007), the initial FoS of the slope is determined by
the slope geometry and initial water table while the actual rainfall-induced failure conditions are
greatly influenced by rainfall characteristics and soil properties. The nature of the rainfall, such as
intensity, duration, spatial distribution and antecedent characteristics, significantly influences the
occurrence of rainfall-induced landslides by affecting the pore water pressure distribution and
increases the self-weight of the slope materials (Abramson et al., 2002; Rahardjo et al., 2008; Rahimi
et al., 2011; Zhang et al., 2011; Mukhlisin and Taha, 2012). Based on several studies performed in
Asia, Lumb (1975) and Brand (1984) indicated that antecedent rainfall has a negligible effect on local
rainfall-induced landslides. They concluded that rainfall intensity and duration had the most
profound influence on the slope failure due to localised high conductivity soils. Later, Rahardjo et al.
(2001, 2008) and Rahimi et al. (2011) undertook a series of more detailed studies along the Sieve
River in Italy in relation to rainfall associated with soils of different conductivities. Their work showed
that the rate of decrease in FoS, the time corresponding to the minimum FoS and the value of the

Page 34 of 77
Riverbank Collapse Task 5 Report

minimum FoS are all controlled by the rainfall distribution. In comparison, rainfall can induce up to a
45% reduction in the FoS of soil slopes with lower fines content and higher saturated permeability
(Ksat ≥ 10-5 m/s) than those with a high fines content and low permeability (K sat ≤ 10-6 m/s) (Rahimi
et al. 2011). These geotechnical properties greatly influence the behaviour of rainfall-triggered slope
failures because they affect rainwater seepage and infiltration. Rahardjo et al. (2007) examined the
relationship between Ksat, soil suction, FoS and the magnitude of rainfall. Their work found that,
under modest rainfall intensity (10 mm/h), soils with Ksat = 10-6 m/s were associated with the lowest
FoS, followed by soil with Ksat = 10-5 m/s. In contrast, under relatively intense rainfall (greater than
200 mm/h), soils with Ksat = 10-5 m/s were associated with the lowest FoS.

River level fluctuation has been shown to influence riverbank stability in two important ways: (i) its
effect on reducing negative pore water pressure and, hence, its consequent reduction in soil strength,
and (ii) the hydrostatic pressure it applies to stabilizing the riverbank (Casagli et al., 1999; Green, 1999;
Rinaldi and Casagli, 1999). Due to the limited models available at the time, studies in the 1980s
typically proposed simple hypotheses on pore water pressure conditions of the riverbank (dry or total
saturated) and adopted relatively simple solutions for slab failures (Hooke, 1979; Higgins, 1980;
Springer, 1981; Thorne, 1982). Later in the 1990s and 2000s, with developments in unsaturated soil
mechanics theory, the effect of pore water pressure on unsaturated riverbanks and confining
pressure became more widely accepted and included into drawdown analysis and research of
riverbank stability (Casagli et al., 1999; Green, 1999; Pauls et al., 1999; Rinaldi and Casagli, 1999; Lane
and Griffiths, 2000; Rinaldi et al., 2004; Zhang et al., 2005; Berilgen, 2007; Jia et al., 2009; Li et al.,
2011; Yang et al., 2010). It is generally accepted that when rivers experience an initial high-flow
period, the riverbanks are stable due to the supportive effect of the hydrostatic pressure of the water.
However, the processes of erosion and soil saturation during high flow events weaken many parts of
the bank by undermining it and reducing the effective strength, respectively (Twidale, 1964; Hooke,
1979; Springer, 1981; Thorne, 1982; Thorne et al., 1997). Berilgen (2007) indicated that the stability
of a submerged slope during drawdown greatly depends on the rate of pore water drainage. While
during initial low-flow periods, the matric suction (the suction due to capillary action and water
surface tension) occasionally allows the riverbank remain stable at steep angles (Casagli et al., 1999).
However, subsequent rainfall increases the dead weight of the bank material and reduces the matric
suction which might be sufficient trigger a mass failure (Rinaldi et al., 2004).

A very limited number of studies modelled the coupling of climatic factors and river level fluctuations.
Casagli et al. (1999) used tensionmeters, piezometers and a rain gauge on the Sieve River to monitor
the matric suction evolution and riverbank stability in a semi-arid climate with daily river flow changes
over a 16-month period. Later, based on these monitoring works, transient modelling of a drawdown
failure, which occurred on December 14, 1996, was carried out by Rinaldi et al. (1999). In the
transient model, the research period (December 13–18, 1996) was divided into 24 time steps to

Page 35 of 77
Riverbank Collapse Task 5 Report

examine the behaviour of the riverbank under different rainfall and flow events. It was indicated that
the minimum FoS always occurred after the peak level of the Sieve River, and no later than 5.5 hours
after. The result suggests that riverbank collapse on Sieve River is dominated by river level
fluctuations, primarily due to the reduction in the stabilizing influence of hydrostatic pressures due to
river level drawdown, and marginally due to rainfall. This conclusion is in agreement with the earlier
work of Twidale (1964), Springer (1981), Thorne (1982, 1988), Thorne et al. (1997) and others.

The section aims to: (i) model the riverbank collapse incident, in both 2D and 3D, which occurred at
Long Island Marina, Murray Bridge, South Australia, on February 4, 2009; (ii) examine the influence
and sensitivity of river level fluctuations and climatic factors on riverbank stability; and (iii) determine
the dominant triggers affecting collapse. Specifically, the study implements a transient water model
in both two and three dimensions to evaluate the FoS of the riverbank by determining the temporal
distribution of pore water pressure in the riverbank and simulating its dynamic response to river level
fluctuations and rainfall. Due to the relatively low width-to-height ratio of the riverbank collapsed
region, as indicated by bathymetric surveys, a 3D analysis is also performed to complement and
validate the date and extent of the collapse and to compare this with the results derived from the 2D
model. Details of the Long Island Marina site are presented in the following section.

4.2 Study Area


The study area is located along the lower River Murray adjacent to the Long Island Marina, Murray
Bridge, South Australia, as shown in Figure 23(a). From the DEWNR inventory, several incidents were
recorded at Long Island Marina between 2008 and 2010. These included 5 major riverbank collapses,
an identified significant tension crack, two riparian trees leaning into the river, and one levee bank
related incident (Section 2). On February 4, 2009 a significant riverbank collapse incident occurred
which involved a 100 m stretch of riverbank [Region (iii) in Figure 23(b)], which failed without warning.
Several large trees, three unoccupied vehicles and an estimated 70,000 m3 of bank material collapsed
into the river (SKM 2011). The bathymetry, as shown in Figure 23(b), illustrates the scale of a series of
collapses, which occurred in the vicinity of Long Island Marina.

As mentioned earlier, a site investigation was performed in October and November 2009 at the Long
Island Marina site by SKM (2010b) in order to obtain geotechnical information related to the collapse.
Details of the investigation and geotechnical characteristics are presented later.

Page 36 of 77
Riverbank Collapse Task 5 Report

(a) Location of study area.

(b) Aerial bathymetry image at Long Island Marina, Murray Bridge showing the contours of the
riverbed and the collapse of 70,000 m3 of material into the River Murray (Source: Beal et al., 2010).

(c) 2D cross-section showing river level on the date of collapse.

Figure 23: Details of the Long Island Marina site.

Page 37 of 77
Riverbank Collapse Task 5 Report

4.3 Methodology
In order to investigate the factors influencing the Long Island Marina riverbank collapse, a transient
slope stability model was developed in both 2D and 3D. The model incorporates riverbank geometry,
geotechnical properties, river level variation, and rainfall and evaporation. The implementation of
each of these into the transient model is outlined in the sub-sections that follow.

SVFlux (SoilVision 2009a) has again been used to undertake finite element analyses of saturated and
unsaturated groundwater flows, in conjunction with SVSlope (SoilVision 2009b), which performs limit
equilibrium slope stability calculations, in this case, in both 2D and 3D. Specifically, SVFlux has been
employed to model the distribution of pore water pressure against time subject to different
circumstances of rainfall, evaporation and river level fluctuations. These have subsequently been
imported into SVSlope to examine the stability of the riverbank at each time-step. The limit
equilibrium method (Bishop’s method of slices) is employed to perform the limit equilibrium slope
stability calculations. The sliding surfaces are determined using the grid and tangent and slope limit
methods, and for the 3D model, reference is made to the actual dimensions and location of the
collapsed region. In order to determine the most appropriate soil properties for the high plasticity,
Silty Clay layer, as explained later, and to accommodate variability of soil properties, back-analyses
are again performed to obtain the closest match between the predicted and actual date of failure (in
the 2D analysis) and the dimensions of the collapsed region (3D analysis).

As mentioned above, in order to determine the riverbank geometries associated with the Long Island
Marina collapse, the GIS framework is adopted using topographic information obtained from the DEM;
the collapsed regions are examined by visual interpretation of high resolution aerial images; and the
dimensions of the simulated collapsed regions are validated against these high-resolution aerial
images.

As mentioned previously, the Long Island Marina riverbank failure is examined in both two and three
dimensions. It is commonplace in the assessment of slope stability for 2D cross-sections to be
adopted (Raharjo et al., 2001, 2008; Rahimi et al. 2011). Stark (2003) undertook a study where he
compared several 2D and 3D analyses and concluded that, if the width-to-height ratio of the collapsed
region is larger than 6, the difference in the factors of safety obtained from the 2D and 3D analyses is
marginal. In the Long Island Marina riverbank collapse, as shown by Region (i) in Figure 23(b) from
the bathymetric data and Figure 23(c) from the cross-sectional model, the width-to-height ratio is
found to be equal to 5.2 (78 m in width and 15 m in height). Hence, 3D modelling is beneficial in this
study by affording an additional analytical technique to validate the volume of the collapsed region
and to compare the FoS time history with that obtained from the 2D model.

Page 38 of 77
Riverbank Collapse Task 5 Report

In order to establish accurate riverbank geometries, ArcGIS is again used, incorporating the bilinear
interpolation method based on elevation extraction. As shown in Figure 24(a) and (b), the 2D cross-
sectional model is initially obtained as a polyline comprised of 401 points with a 10 m interval
between neighbouring points. The location of the cross-section used in the 2D analysis is selected to
coincide with the recorded collapse, as indicated in the DEWNR riverbank collapse inventory and the
bathymetric data [Region (i)], as shown in Figure 23(b) and Figure 24(b). Based on the bilinear
interpolation method, the elevation values were extracted from two DEMs (a 1 m resolution model
acquired in 2008 and b 0.2 m resolution model acquired in 2010) and assigned to 401 points, with
singular points avoided. This approach yields accurate and high-resolution cross-sections, suitable for
subsequent importing into SVFlux and SVSlope.

By utilising the high-resolution DEMs, the topography of the riverbank region of interest is extracted
using custom masks within ArcGIS, as shown in Figure 24(b) for the purpose of reproducing the
riverbank model in three dimensions. Due to the nature of DEM data, the elevation values that are
extracted are irregularly distributed. To resolve this issue, the inverse distance weighting method
(IDW) is used to interpolate a new raster based on the former irregularly distributed elevation values.
As shown in Figure 24(c), the geometry of the riverbank was established in 3D in SVFlux3D from the
updated elevation values extracted from the new raster.

The dimensions of the collapsed regions of the riverbank need to be identified with a relatively high
level of accuracy in order to facilitate the proposed modelling. In this study, this has been achieved
by undertaking a visual interpretation of the high-resolution aerial images taken on different dates
(Section 2). By comparing the 0.5 m resolution aerial photos acquired in March 2008 [Figure 25(a)]
with the 0.2 m resolution photos acquired in May 2010 [Figure 25(c)], the collapsed Region A has
been accurately identified and vectorised. Comparing the collapsed regions (i), (ii) and (iii) shown in
Figure 23(b), Region A [shown by the dotted region in Figure 25(c)] is the section above the river level.
To assist with the visual interpretation of the collapsed region, LIDAR DEMs obtained in 2008 [Figure
25(b)] and 2010 [Figure 25(d)] have been employed, especially to verify the elevation of the riverbank
prior to the collapse with river levels at the same location subsequent to the collapse. For example,
location A' [shown by the red dot in Figure 25(c)], where the riverbank was at 0.6 m AHD prior to the
collapse, is compared with the river level at location B', which is at −0.45 m AHD subsequent to the
failure. By adopting this comparative process, the identified region A is confirmed as having collapsed
rather than simply having been submerged.

The geotechnical model is developed from the results of a geotechnical investigation performed at
Long Island Marina by SKM (2010b), which included the drilling of two boreholes to a depth of 20 m,
12 piezocone tests (CPTUs) with pore water pressure (u) measurements, and a series of vane shear
and laboratory tests (Figure 26 to Figure 29).

Page 39 of 77
Riverbank Collapse Task 5 Report

(a) Elevation extraction from the 2008 (b) Elevation extraction from the 2010 LIDAR
LIDAR DEM. DEM, with river level extracted.

(c) 3D topographical model.


Figure 24: Riverbank geometry definition.

In accordance with the borehole logs from SR-BH1 and SR-BH2 (SKM 2010b), the bank profile at Long
Island Marina has previously been modelled by SKM (2010a) and Coffey (2012) using three separate
soil layers. Specifically, in the SKM and Coffey models, as summarised in Table 5, an effective stress
analysis was performed in the Fill and SC layer using the Mohr-Coulomb failure criterion, while a total
stress analysis was adopted in the CH layer using a depth-dependent, undrained soil model. The
adopted depth-dependent, undrained soil model incorporates linearly-increasing cohesion with depth,
with ctop quantifying the cohesion (kPa) at the upper layer boundary, c ratio representing the gradient of

Page 40 of 77
Riverbank Collapse Task 5 Report

(a) Aerial photo in 2008. (b) LIDAR DEM before collapse.

(c) Aerial photo in 2010 with collapsed region (d) LIDAR DEM after collapse.
highlighted.
Figure 25: Example of adopted visual interpretation process on high-resolution, aerial images within ArcGIS.

increasing cohesion with depth and capped at a maximum value of c max (SKM 2010a). However, in
order to accommodate the effects of positive and negative pore water pressures in unsaturated
riverbanks, an unsaturated stress analysis is performed in the Fill and CH layer for this paper, which
will be discussed later. As summarised in Table 5, the values recommended by SKM (2010a) and
Coffey (2012) were inferred from laboratory and in situ test results obtained from their respective
geotechnical investigations.

Table 5: Soil parameters for stability assessment.

Undrained parameters for CH


Elevation c' γ Ksat θsat
Layer φ' layer from:
(m AHD) (kPa) (kN/m3) (m/day) (%)
SKM (kPa) Coffey (kPa)
Silty/Clayey
Sand 1 to 0 c'=2 ± 2 28° ± 2° 18 ± 1 – – 13.51 52.7
(SM/SC)

cu-top = 10 ± 5 cu-top = 5.5


Silty Clay
0 to –20 0 27° 16 ± 1 cu-ratio = 1.25 cu-ratio = 1.25 9.9x10–5 63.2
(CH)
cu-max = 25 ± 5 cu-max = 25 ± 5

Clayey
Sand/ Sandy > –20 c'=2 ± 2 30° ± 2° 17 ± 1 – – 0.187 54.6
Clay (SC/CL)

Page 41 of 77
Riverbank Collapse Task 5 Report

(a) Moisture content and plasticity index profile, (b) Dry density profile and undrained shear
Type-A represents SM/SC, Type-B represents CH strength from CPTu.
and Type-C represents SC/CL.
Figure 26: Geotechnical profiles based on soil samples taken from SR-BH1 and SR-CPTu6s at Long Island Marina.

Figure 27: Particle size distributions based on the soil samples from four different depths in borehole SR-BH1.

Page 42 of 77
Riverbank Collapse Task 5 Report

In order to determine accurately the variation of total soil suction with time and moisture, a soil
water characteristic curve (SWCC) is adopted in the finite element method analyses for unsaturated
soils within SVFlux. The Zapata et al. (2000) method is used to estimate the SWCC in each of the soil
layers from the particle size distributions (PSDs) shown in Figure 27. The weighted Plasticity Index
(wPI) is used within the Zapata method and is defined as an indicator for water adsorption and
retention in soil particles with different surface areas (Zapata et al. 2000). As stated by Zapata at al.
(2000), a proportional relationship exists between the equilibrium soil suction at a given degree of
saturation and the specific surface area of the soil. Compared with the plasticity index (IP), wPI is
considered as a better indicator, especially for the soils that have a high IP but are associated with
relatively small surface areas (Zapata et al., 2000). As shown in Table 6, for soil with a high plasticity
(wPI > 0), such as the Silty Clay (CH), the Fredlund et al. (1995) parameters a f, bf, cf and hf are
calculated based on the relationships shown in Table 6 to estimate the SWCC for each of the three
layers (Figure 28). In contrast, for granular soils with wPI = 0 (i.e. non-plastic soils), the parameter D60,
which represents the grain diameter corresponding to 60% passing by weight from the PSD, is
employed.

Figure 28: Estimated SWCCs for the three soil layers at Long Island Marina using the Fredlund and Xing (1994)
estimation method.

The saturated hydraulic conductivity, Ksat, is derived from a CPTu pore pressure dissipation test
performed in the CH layer at 5 m depth at the Murray Bridge site [Figure 13(b)]. Based on
relationships between measured excess pore pressures and the saturated hydraulic conductivity
(Burns and Mayne, 1998) for the clay layer, Ksat is found to equal approximately 9.86 × 10–5 m/day, as
shown in Table 5. The values of Ksat for the Fill and SC layers are obtained from literature (Table 5) as
no measurements are available. In unsaturated soil, the unsaturated hydraulic conductivity, K, varies
with respect to matric suction, and is calculated indirectly from the Fredlund and Xing estimation

Page 43 of 77
Riverbank Collapse Task 5 Report

Table 6: Equations used to calculate Fredlund and Xing (1994) SWCC fitting parameters based on the soil grain
size distribution.

Soil with high plasticity Non-plastic soil

𝑤𝑃𝐼 > 0 𝑤𝑃𝐼 = 0

𝑎𝑓 = 1.14𝑎 − 0.5
𝑎𝑓 = 32.825 𝑙𝑛(𝑤𝑃𝐼) + 32.438
4.34
𝑎 = −2.79 − 14.1 𝑙𝑜𝑔(𝐷20 ) − 1.9 ∗ 10−6 𝑃200 + 7 𝑙𝑜𝑔(𝐷30 ) + 0.055𝐷100

𝑏𝑓 = 0.936𝑏 − 3.8

𝑏𝑓 = 1.42(𝑤𝑃𝐼)−0.3185 𝐷90
𝑏 = (5.39 − 0.29 𝑙𝑛 (𝑃200 ⁄𝐷 ) + 3𝐷00.57
10
1.19 30
+ 0.0021𝑃200 )( ⁄𝑙𝑜𝑔(𝐷 ) − 𝑙𝑜𝑔(𝐷 ))0.1
90 60

𝑐𝑓 = 0.26𝑒 0.758𝑐 + 1.4𝐷10


𝑐𝑓 = −0.2154 𝑙𝑛(𝑤𝑃𝐼) + 0.7145
𝑐 = 𝑙𝑜𝑔(20⁄𝑙𝑜𝑔(𝐷 ) − 𝑙𝑜𝑔(𝐷 ))1.15 − 1 + 1⁄𝑏
30 10 𝑓

ℎ𝑟 = 500 ℎ𝑟 = 100

4(𝑙𝑜𝑔(𝐷90 )−𝑙𝑜𝑔(𝐷60 ))⁄


𝐷100 = 10 3+𝑙𝑜𝑔(𝐷60 )
𝑤𝑃𝐼 = 𝐼𝑃 × 𝑃200 ⁄100
−3(𝑙𝑜𝑔(𝐷30 )−𝑙𝑜𝑔(𝐷90 ))⁄
𝐷0 = 10 2+𝑙𝑜𝑔(𝐷30 )

Note: 𝑃200 is the percentage of soil passing the US standard sieve #200. IP is the plasticity index as shown in
Figure 26(a); and 𝐷% is the grain diameter related to the percentage of passing in mm.

associated with Ksat (Fredlund et al., 1995). Adopting Equation (3), the saturated volumetric water
content, θsat, is determined by the following equation and based on the tests from borehole soil
samples (Table 5).

ρ
θsat = 1 − ⁄G ρ (1 + w) (3)
s w

where ρ is the dry density; ρw is the density of water at 4°C and w is the moisture content.

In order to model the influence of rainfall on riverbank stability it is necessary to estimate the runoff
coefficient so that infiltration of rainfall into the riverbank can be evaluated. It is well understood that
the runoff coefficient is a function of hydraulic conductivity and thickness of the near-surface soil
layers, the type of surface vegetation and land use. Due to the high rate of evapotranspiration,
temporal and spatial variability in rainfall intensity and frequency, and the generally flat topography
across most of South Australia (National Water Commission, 2005) the runoff coefficient is relatively
small. Hence, based on the Australian Rainfall and Runoff (Institution of Engineers Australia, 1987) a
runoff coefficient of 6 is established for Murray Bridge for an average recurrence interval of 10 years.

Page 44 of 77
Riverbank Collapse Task 5 Report

The River Murray level data used in the analyses were obtained from Murray Darling Basin Authority
for Murray Bridge at observation station (A4261162) incorporating data from No. 1 Pump Station
(A4260522) and (A4261003) (Murray Darling Basin Authority, 2014). From these data, the river levels
at Murray Bridge had remained relatively constant between 0.5 to 1 m AHD until the end of 2006,
where they started to fall continuously and significantly to −1.5 m AHD until September 2009. They
then gradually returned to the pre-2006 river levels. As shown in Figure 29, the study period from
May 1, 2008 to February 28, 2009 (304 days) is selected during this period of low river levels to model
the Long Island Marina riverbank collapse with the combined influence of river level fluctuation,
rainfall and evaporation. For simplicity, the water level in the embayment area is assumed to be
constant at 1 m AHD.

Figure 29: Daily river levels, daily rainfall and daily mean temperature from May 1, 2008 to February 28, 2009
at Long Island Marina.

As summarised in Figure 29, mean daily rainfall, as well as mean daily temperature, at the Long Island
Marina site were collected from the Bureau of Meteorology (2014), in addition to evaporation and
humidity, in order to provide climatic data for the riverbank stability modelling.

Within the SVFlux2D framework, the riverbank geometry is established from the DEM and ArcGIS, as
mentioned above and shown in Figure 30. The boundary conditions are established to represent the
actual site circumstances as: boundary AH was set to ‘climate’ which represents the combined effects
of precipitation and evaporation over the 304 day study period; boundary HG is established as ‘head
data review’ to account for the effects of river level variation over this period; and boundaries BH and

Page 45 of 77
Riverbank Collapse Task 5 Report

CF are set to the ‘no boundary’ condition, which indicates a barrier free condition for water flux; and
the remainder are set to ‘zero flux’ to simulate no water flux through these boundaries.

(a) 2D cross-sectional model.

93
m 51 m

(b) 3D slope stability model.


Figure 30: Results of (a) 2D (Day 282: February 6, 2008) and (b) 3D (Day 287: February 10, 2008) riverbank
stability analyses of Long Island Marina site.

Page 46 of 77
Riverbank Collapse Task 5 Report

In the SVFlux3D model, in order to facilitate modelling pore water pressure variations due to river
level fluctuation and climatic influences, the riverbank is divided into two zones, as shown in Figure
24(c). The climatic data of rainfall, evaporation, air temperature and humidity were assigned the
‘climate’ boundary condition to the ground surface of Zone I, while the river level fluctuation was
applied as a ground surface boundary condition in Zone II as ‘head data review’; the sidewall
boundary between Zones I and II was assigned the ‘no boundary condition’ to permit flux; while the
remaining three sidewalls associated with both zones were assigned ‘zero flux’ to simulate no water
flux.

4.4 Results
The results of the 2D back analyses, using SVFlux2D and SVSlope2D, and those in 3D, using SVFlux3D
and SVSlope3D, are presented as follow. The resulting pore water pressure and FoS time histories are
examined with reference to river level fluctuation and climatic influences. As mentioned above,
SVFlux is used to model the variation in pore water pressure (PWP) as a consequence of rainfall,
evaporation and river level fluctuations based on the transient seepage model. As shown in Figure 30,
5 nodes were selected to monitor the PWP variation above the slip surface at locations T1, T2, T3, T4
and T5.

Figure 31 shows the evolution of PWPs as obtained from SVFlux2D at locations T1 – T5, with (solid
lines) and without (dashed and dotted lines) the effects of evaporation. As indicated, the entire study
period of 304 days is divided into two parts: Part A includes late autumn and winter (the rainy season
in South Australia); and Part B the subsequent spring and summer seasons, which are warmer and
drier.

Horizontally, the daily PWP profiles showed that although T1, T4 and T5 are at the same elevation,
each has different pore water pressures due to the head variations between embayment and river
level, but they exhibit a very similar trend throughout the 304-day period. This is also true when
comparing T2 and T3. Specifically, at locations T1, T4 and T5, the profiles of PWP increased in late
autumn and winter 2008 corresponding to rising river level, and then reduced in spring and summer
2008 as the river level dropped. The fluctuations in PWP profiles are caused by the climatic factors
(i.e. rainfall, temperature and evaporation). The PWP at T2 and T3 remain unchanged during autumn
and winter, but recorded increased in spring and summer, probably due to larger head variations
between embayment and river level.

For T1, T4 and T5, there is very little effects of the evaporation on the evolution of PWPs evaporation
in autumn and winter, but the effects are more pronounced in spring and summer, as the lines (with

Page 47 of 77
Riverbank Collapse Task 5 Report

Figure 31: Evolution of pore water pressure at 5 selected nodes through the entire research period accounting
for, and without, evaporation.

and without the influence of evaporation) have larger divergence at the beginning of spring (early
September 2008). There is no effect of evaporation on T2 and T3 as soils at that depth are shielded
from the elements. Therefore, the sensitivity to the combined effects of rainfall, evaporation and
lateral bank seepage, as a consequence of river level fluctuation, diminishes with increasing depth.
For example, as a consequence of the 3-day storm between December 12–14, 2008, there is
immediate change in PWP at T1 than at T2, when comparing the PWPs before and after the storm.

During Part A, as indicated in Figure 31, roughly 164.7 mm rainfall was recorded (70% of that over the
entire 304 day period), the river rose by approximately 0.3 m, and the lower daily temperatures
resulted in modest daily evaporation, there is no to modest changes in PWP in each of the 5 nodes. In
contrast, in Part B, a moderate rainfall (81.5 mm), accompanied by relatively high daily evaporations
and a significant drawdown of river level (0.9 m), contributed to a significant reduction in pore water
pressure. A summary of the PWP distribution is shown in Figure 32, with Figure 32(a) showing the
PWPs during the day with the highest river level (Day 138: 15/09/2008) and Figure 32(b) illustrating
the PWPs during the day with the lowest river level (Day 302: 26/02/2009).

As mentioned previously, deep-seated, circular slip failures in the soft and very soft clays of Holocene
age were indicated as the dominant bank collapse mechanism along the lower River Murray.
Therefore, back-analyses are performed using the limit equilibrium method (Bishop’s method of slices)
implemented in SVSlope2D and SVSlope3D. Given that the riverbank collapse is recorded to have

Page 48 of 77
Riverbank Collapse Task 5 Report

(a)

(b)
Figure 32: PWP distributions as a result of (a) the highest (Day 138) and (b) lowest (Day 302) river levels.

occurred on February 4, 2009, the back analyses are benchmarked against this date, as well as the
dimensions and extent of failure indicated by the bathymetric survey summarised in Figure 23(b). The
geotechnical properties of the high plasticity, Silty Clay (CH) layer were varied marginally against the
results of the site investigation (Table 5) until a FoS of 1.00 was obtained close to the recorded date of
collapse.

Figure 33 presents the results of the 2D and 3D back analyses in terms of FoS as a function of date.
The nearest boreholes and CPTus suggested an unsaturated Fredlund mode (Fredlund et al., 1996) for
the Fill and CH layers, which is able to accommodate soil suction changes in the unsaturated bank
materials, as shown by Equation (2).

Based on the models adopted and data obtained by SKM (2010b), the following models are
incorporated in the analyses which follow: Fill layer: c' = 0 kPa; φ' = 28°; and CH layer: φ' = 0 kPa; φ' =
22°. The latter model aligns reasonably well with the single CU triaxial test performed by SKM (2010b)

Page 49 of 77
Riverbank Collapse Task 5 Report

(φ' = 27°), noting that only one soil sample was tested and the model adopted in the analyses utilises
an average value of c' throughout the entire CH layer.

2.0

1.8 CRLM 2D Model 3D Model

1.6
Factor of Safety

1.4

1.2

1.0

0.8
6/1/08 7/1/08 8/1/08 9/1/08 10/1/08 11/1/08 12/1/08 1/1/09 2/1/09

Date (dd/mm/yy)

Figure 33: Factors of safety from the 2D, 3D and CRLM models.

The results of the 2D analyses shown by the dashed line in Figure 33 are extremely encouraging.
Firstly, the predicted date of failure is Day 282 (February 6, 2009) compares very favourably with the
recorded date of failure February 4, 2009 (Day 280). Secondly, the FoS is consistently above 1.00
from the beginning of the analysis period until the predicted date of failure. Finally, the predicted slip
surface is shown in Figure 30(a). The width of the collapsed region from the crest to the riverbank
line is 19.8 m, which again compares extremely favourably with the actual width of 20 m determined
from aerial photos [Figure 25(b)].

As mentioned above, a 3D slope stability analysis is performed to complement and enhance the 2D
riverbank slope stability analyses. Generally, 2D riverbank slope stability analysis, which assumes that
the failure surface is infinitely wide, is suitable for stability estimation or assessment over a relatively
extensive region, such as adopting a large number of 2D cross-sections, spaced at appropriate
intervals, to assess the stability of a large extent of the river. However, 3D riverbank slope stability

Page 50 of 77
Riverbank Collapse Task 5 Report

analysis is generally more accurate for collapse back-analysis over relatively small regions where the
dimensions or volume of the collapse zone are known (Stark and Eid, 1998). As shown in Figure 23(b),
the bathymetric data clearly shows three, what appear to be, separate failures, identified as Regions (i)
– (iii). From the DEWNR database of recorded riverbank collapse incidents, it is clear that Region (iii)
has largest collapsed volume of soil, and hence attention is directed solely to this failure. The
modelling of Regions (i) and (ii) is beyond the scope of this paper. The failure of Region (i),
approximate a year prior to Region (iii), has some effect on the pre-collapse topography associated
with Regions (ii) and (iii) and, given the limited extent of available observational and survey data,
presents significant challenges and uncertainties, therefore, the effects are not quantified and
ignored in this analysis. Figure 33 also shows, by means of a solid line, the results of the 3D back-
analysis of the riverbank. As for the 3D analysis, an unsaturated Fredlund model is again adopted
with the following properties: c' = 0 kPa; φ' = 20°. As can be seen, this 3D model omits the Fill layer
due to insufficient information and large computation resources required (additional finite elements)
to incorporate this thin layer of Fill into the model. The omission represents a modest reduction in
soil weight and cohesion when compared with the 2D model and also the reduction in φ' from 24° to
22° in the 3D model.

As shown in Figure 33, the reductions in FoS can be observed in the 3D model, which likely relate to
the reduction in the daily river level and the effects of large rainfall events. As can be observed, river
level fluctuation dominates the likelihood of FoS variation in both 2D and 3D. Compared with the 2D
model, the 3D model FoS appears to be less sensitive to river level variation, however a significant
drawdown may result in a significant reduction in FoS. Again, the results represent extremely good
correlation between the predictions and observations. Firstly, the predicted date of failure Day 291
(February 15, 2009) aligns well with the recorded date of failure February 4, 2009 (Day 280). It is
noted that the extra loads applied by the three vehicles that slipped into the river are not included in
the model. The FoS is predicted to be close to unity on February 4, 2009, and coincidentally, the
additional vehicle weights might well be sufficient for failure to occur on that day. Secondly, the FoS
is again consistently above the unity from the beginning of the analysis period until close to the
predicted date of failure. Finally, the predicted volume of the slip surface is in Figure 9(b) and the
volume and surface area of the collapse region is comparable to the results of the bathymetric survey.

In order to understand better the relationship between rainfall, river level fluctuations and riverbank
stability, additional 2D analyses are performed using SVFlux and SVSlope. The first series of analyses
examines the situation where the river level remains static while rainfall continues to vary in
accordance with the historical record. The second series examines the situation of extreme rainfall
events. Figure 33 and Figure 34 present the results of analyses adopting a constant river level model
(CRLM) with a static level at –0.521 m AHD for the entire 304 day study period. This level coincides
with the initial river level associated with the actual historical model (HM) examined in the previous

Page 51 of 77
Riverbank Collapse Task 5 Report

analysis. As can be seen from Figure 33, compared with the HM 2D and 3D factors of safety, the
CRLM FoS generally remains constant and consistently greater than the unity, with several minor
reductions as a result of intensive rainfall events.

Figure 34(a) presents the time histories of FoS and PWP in both the HM and CRLM in mid-December
2008, when the most intense storm of the 304 study day period resulted in 29.8 mm of rainfall in 3
days. It can be observed that the PWPs associated with the CRLM at location T1 (refer to Figure 31(a)
are considerably lower than those derived from the HM. This is because the lower river level (–0.521
m AHD) creates a depressed groundwater table resulting in decreased PWPs. It can also be observed
in Figure 34(a), that the riverbank is less stable in the HM than the CRLM on 11 December due to the
lower river level, but more stable in the HM than in the CRLM after the storm. The river levels during
these 5 days (10–14 December) were –0.439, –0.486, –0.421, –0.278 and –0.344 m AHD, respectively.
It can also be seen clearly that the CLRM FoS is generally constant throughout mid-December 2008,
albeit with a negligible reduction after the storm. This demonstrates that daily river level fluctuation
has an observable influence on the FoS, while the effect of daily rainfall is modest. In contrast,
relatively minor precipitation was recorded (0.8 mm) during the intervening period leading up to the
collapse on 4 February 2009. As shown in Figure 13(b), the HM FoS continues to decline until collapse
is predicted on 6 February 2009. During this period the CRLM FoS remains constant and consistently
greater than the unity.

Further examining the influence of rainfall on riverbank collapse, an extreme rainfall scenario is
modelled. The most significant storm on record at Murray Bridge occurred in January 1941 when
189.6 mm of rain fell over a period of 5 days. This is more than 6 times greater than the storm
examined above.

(a) HM under the storm occurred in Dec. 2008. (b) MRM with high daily river levels (approx.
–0.2 m AHD).

Page 52 of 77
Riverbank Collapse Task 5 Report

(c) MRM with medium daily river levels (approx. (d) MRM with low daily river levels (approx.
–0.5 m AHD). –0.8 m AHD).
presents the results of the magnified rainfall model (MRM), which incorporates the January 1941
storm. Four different scenarios are presented: Figure 35(a) shows the results of the HM for
benchmarking purposes; whereas Figure 35(b) – (d) present the results of the MRM with high
(–0.2 m AHD in September 2008), medium (–0.5 m AHD in May 2008) and low (–0.8 m AHD January
2009) river levels, respectively.

Figure 35(a) shows the behaviour of the HM as a result of the historical storm that occurred in
December 2008. The FoS remains consistently above unity with daily river levels fluctuating around
–0.5 m AHD. In contrast, the FoS resulting from the MRM with high and medium daily river levels, as
shown in Figure 35(b) and (c) respectively, reduces significantly with collapse predicted. The extreme
storm results in an approximate increase of 30 kPa in PWP, and consequently triggers a rapid and
significant decrease in FoS in both scenarios.

Page 53 of 77
Riverbank Collapse Task 5 Report

(a) Storm in mid-Dec. 2008.

(b) Scenario of the riverbank collapse at Long Island Marina on 4 Feb. 2009.
Figure 34: Factors of safety for historical model (HM) and constant river stage model (CRLM) in two scenarios.

Page 54 of 77
Riverbank Collapse Task 5 Report

(a) HM under the storm occurred in Dec. 2008. (b) MRM with high daily river levels (approx.
–0.2 m AHD).

(c) MRM with medium daily river levels (approx. (d) MRM with low daily river levels (approx.
–0.5 m AHD). –0.8 m AHD).
Figure 35: Magnified rainfall model (MRM) under different river level scenarios.

Figure 35(d) shows a similar outcome to that of the HM in Figure 35(a), where the FoS remains
relatively consistent above unity, in the former as a result of the MRM with low daily river levels.
After the extreme rainfall event, the change in PWP at T1 is 20 kPa, which is roughly 10 kPa less than
that observed in the high and medium river level scenarios in Figure 35(b) and (c).

A sudden or rapid drawdown scenario has also been examined based on the HM. The river level was
modelled to decrease from the highest point (–0.005 m AHD, Day 138 of HM) to the lowest (–1.002 m
AHD, Day 302 of HM) within 2 days (drawdown ratio = 0.5 m/day) and 5 days (drawdown ratio =
0.2 m/day). The results show that, the greatest drawdown ratio (0.5 m/day) is associated with the
lowest FoS (0.88), followed by drawdown ratio = 0.2 m/day (FoS = 0.914). The HM (164-day
drawdown period with a drawdown ratio = 0.006 m/day) has the highest relative FoS of these three
scenarios, albeit less than unity.

The analyses presented in this section have demonstrated that the river level fluctuations have a far
greater influence on riverbank collapse than rainfall. However, if an extreme rainfall event coincides

Page 55 of 77
Riverbank Collapse Task 5 Report

with a medium (–0.5 m AHD) to high (–0.2 m AHD) river level, collapse is likely to occur. Whilst not a
key objective of the present study, a series of additional analyses have also been performed to
examine the influence of evaporation on FoS. The results, however, demonstrate that evaporation
has a marginal effect on FoS.

4.5 Summary
In summary, this section has sought to: (i) model the riverbank collapse incident, in both 2D and 3D,
which occurred at Long Island Marina, Murray Bridge, South Australia, on February 4, 2009;
(ii) examine the influence and sensitivity of river level fluctuations and climatic factors on riverbank
stability; and (iii) determine the dominant triggers affecting collapse. The modelling has been
undertaken using the limit equilibrium method in 2D and 3D using SVFlux and SVSlope. To facilitate
the stability analyses, a GIS framework has been adopted, which involved: examination of the
recorded collapsed regions by visual interpretation of high-resolution aerial images; obtaining
topographic information of the site from a digital elevation model; and calculation of the dimensions
of the collapsed regions from high-resolution aerial images. Back-analyses have been performed
using geotechnical data obtained from a site investigation incorporating both in situ and laboratory
testing and validated against data from the recorded collapse.

This study has shown that:


(a) Both the adopted process of 2D and 3D stability modelling yielded excellent predictions of the
collapse when compared against the recorded date of collapse and dimensions of the failed
region. The adopted geotechnical model parameters align with those derived from the site
investigation.
(b) The integration of GIS with transient unsaturated soil modelling has proved to be an effective
tool for accurately predicting riverbank stability.
(c) River fluctuations, rather than climatic factors, dominate the likelihood of riverbank collapse
along the Lower River Murray.
(d) Sudden or rapid drawdown is associated with a lower FoS compared with the historical model.
(e) However, extreme rainfall events, coinciding with medium to high river levels, are also likely to
trigger riverbank collapse.

Page 56 of 77
Riverbank Collapse Task 5 Report

5 Identifying the Areas Susceptible to Riverbank Instability

5.1 Introduction
In the Task 4 report entitled Mechanisms and Processes of the Millennium Drought Riverbank Failures
(Hubble and De Carli, 2015) it was stated that the majority of the larger bank failure features are
associated with deep holes that have been eroded into the channel floor. These deep holes are
formed due to either: (a) bedrock constriction and pronounced narrowing of the channel cross-
section; or (b) large outcrops of bedrock which protrude up from the floor of the channel (Hubble and
De Carli, 2015). This has generated erosive flow patterns during periods of higher flow that have
eroded or scoured deep holes in the channel and over-steepened the channel margins and banks.
Riverbanks in reaches of the channel affected in this way are less stable than the banks present
elsewhere along the Lower Murray where the channel profile is wider and shallower (Hubble and De
Carli, 2015).

Hubble and De Carli (2015) also identified the failure prone sites which possess the geomorphologic
features of deep holes through bathymetric studies, namely: Thiele Reserve, Bells Reserve,
Whitesands, Sturt Reserve, Long Island Marina, Woodlane Reserve and East Front Road near
Younghusband. Over-steepened riverbanks and local anthropogenic modification of the banks (e.g.
placement of fill or construction of the embankments near the waterline), combined with low pool-
levels, are identified as the main causes or factors affecting riverbank stability. Numerical modelling
was undertaken by Hubble and De Carli (2015) to study the effects of the aforementioned factors on
the FoS of the riverbanks at the Thiele Reserve.

This section of the report will expand the numerical studies to determine the FoS of the riverbanks at
Thiele Reserve when the river level is at normal pool level, i.e. 0 m AHD. Geographic information
system (GIS) modelling and mapping has again been used to facilitate the study. A methodology,
combining both SVSlope and ArcGIS, has been developed to identify the riverbanks along the Lower
River Murray that are susceptible to instability.

5.2 Methodology and Results


As demonstrated in the earlier sections slope failure susceptibility prediction, at medium or large
scales, 
 GIS modelling and mapping greatly enhances and simplifies the analyses. ArcGIS can assist in
locating the deeps holes and over-steepened riverbanks, an example of which is shown in Figure 36.
By extracting the contour lines from the DEM using ArcGIS [Figure 36(b)], and by using its filtering
features, the contour lines at or below –10 m AHD [Figure 36(c)] can be highlighted to identify the
location of the deep holes near Thiele Reserve. The same process can be applied to the Lower River
Murray and the results are shown in Figures Figure 36 to 39.

Page 57 of 77
Riverbank Collapse Task 5 Report

N (a)

(b)

Avoca Dell

N
Thiele Reserve

Murray Bridge East

(c)
Murray Bridge

Figure 36: Locating deep holes near Thiele Reserve using: (a) bathymetry from Hubble and De Carli (2015)
(areas of bedrock margins are represented by dashed white line); (b) a contour map obtained from ArcGIS
superimposed on satellite imagery from Google Maps; (c) blue contour lines showing elevations that are
–10 m AHD or deeper, whilst the green contour lines represent 0 m AHD.

Page 58 of 77
Riverbank Collapse Task 5 Report
Lock 1

Blanchetown

Blanchetown

Swan Reach

Punyelroo Big Bend


Wongulla

Big Bend

Nildottie

Wongulla
Walker Flat

Walker Flat

Figure 37: Locations of deepened channel beds downstream from Blanchetown to Walker Flat (blue contour lines show elevations that are at –10 m AHD or deeper, whilst the green contour lines are at 0 m AHD).

Page 59 of 77
Riverbank Collapse Task 5 Report

Younghusband Walker Flat


Caloote
Wall Flat

Caurnamont Purnong
Woodlane

Younghusband

Bowhill

Mannum

Mypolonga

Caloote
Caurnamont
Wall Flat Purnong

Woodlane

Mypolonga
Avoca Dell

Murray Bridge East Avoca Dell


Bowhill
Murray Bridge East

Figure 38: Locations of deepened channel beds downstream from Walker Flat to Murray Bridge East (blue contour lines show elevations that are at –10 m AHD or deeper, whilst the green contour lines are at 0 m AHD).

Page 60 of 77
Riverbank Collapse Task 5 Report

Monteith

Thiele Reserve
Murray Bridge East
Murray
Bridge

White Sands

Thiele Reserve

Murray Bridge East Monteith Tailem


Bend

White Sands

Jervois
Sturt Reserve
Woods Point
Long Flat

Long Island
Marina
Murray Bridge
Tailem
Woods
Bend
Point Jervois

Wellington
East
Wellington Wellington
East
Wellington

Figure 39: Locations of deepened channel beds downstream from Murray Bridge East to Wellington (blue contour lines show elevations that are at –10 m AHD or deeper, whilst the green contour lines are at 0 m AHD).

Page 61 of 77
Riverbank Collapse Task 5 Report

The results show that there are a number of deep holes from Blanchetown to Purnong and they are
generally separated by large distances (refer to Figures 37 and Figure 38). Deepened channels are
more common from Bowhill to Younghusband, and from Mannum to Wellington (see Figure 38 and
39). As a case study, the effects of the deep holes on the FoS of the riverbanks near Thiele Reserve
will be examined, along with other contributing factors.

By using ArcGIS, a series of transects has been created along the waterline, at the normal pool level of
0 m AHD near Thiele Reserve (Figure 40), and the elevations along the transects have been extracted
using the bilinear interpolation method at a one-metre interval. This approach yields accurate cross-
sections, suitable for subsequent importing into SVSlope2D.

Thiele Reserve

Murray Bridge East

Figure 40: A series of transects along 0 m AHD contour lines near Thiele Reserve (blue contour lines show
elevations that are at –10 m AHD or deeper, whilst the green contour lines are at 0 m AHD. The red dot
marks the approximate location of the past riverbank instability incident and the dashed orange line
marks the areas of the bedrock margins).

As in the previous sections, SVSlope2D, which performs limit equilibrium slope stability calculations in
2D, has been adopted to study riverbank stability. The geotechnical profile and the parameters are
determined from a combination of field and laboratory testing. A single onshore CPTu has been
undertaken (Jaksa et al., 2015), and the results showed that the subsurface profile encountered at

Page 62 of 77
Riverbank Collapse Task 5 Report

Thiele Reserve is generally comprised of Fill overlying Silty CLAY with a transition occurs at around
0.8 m AHD. The layer of dark-grey, very-soft Silty CLAY with high plasticity fines, is relatively thick,
extending down to –14.3 m AHD (Jaksa et al., 2015). The investigation confirmed the expected
Quaternary aged Alluvial Flat Deposits, as seen on the Geological Survey of South Australia (1962)
Barker map Sheet 1 54-13. A Clayey SAND and Sandy CLAY interbedded layer underlies the Silty CLAY,
extending to the refusal depth. Furthermore a series of consolidated undrained (CU) triaxial tests
were carried out on soil samples collected from Thiele Reserve using the procedures recommended
by Head (1998). Drained parameters c' and ' were found to be 0 kPa and 41 respectively (Jaksa et
al., 2015). A summary of the geotechnical parameters adopted for numerical modelling is presented
in Table 7.

Table 7: Adopted geotechnical parameters for stability calculations.

Layer Elevation (m AHD) c' (kPa) φ' (°) γ (kN/m3)

Fill Ground surface to −0.8 0 28° 18

Silty Clay (CH) −0.8 to −14.3 0 41° 16

Sand > −14.3 0 30° 17

Subsequently, the bank profiles at Thiele Reserve are modelled as three separate, homogenous soil
layers. For simplicity, the soil layers and the transitions zone between the soil layers are assumed to
be horizontal. An effective stress analysis was performed in the Fill, CH and Sands layer using the
Mohr-Coulomb failure criterion. Effective stress analysis was used to assess the long-term stability of
the riverbanks. An example of the soil profile, as well as geotechnical parameters used in the analysis
is shown in Figure 41 and an example of the result is shown in Figure 42.

Stability analyses have been undertaken from Transects No. 4207 to 4242, and also No. 5527 to 5541
using the soil profiles (Fill-Clay-Sand) outlined previously. The results of the analysis are summarised
in Figure 43. Transects No. 5510 to 5526, where the bedrock margin is located (represented by
dashed orange line in Figure 40), have not been analysed. This is because the subsurface of these
transects has different geotechnical profiles than that associated with hard bedrock and slip circle
analysis is unsuitable for this type of subsurface material. The yellows dots shown in Figure 43
represent the locations of the riverbanks that are vulnerable to instability in the long-term and also
the locations of the top of slipe circles, normally at the crest of the riverbanks and very close to the
0 m AHD waterline. The locations of the levees are also shown as thick green continuous lines in
Figure 43. It is noted that, in contrast to the earlier work, the climatic influences, such as rainfall and
evapotranspiration, are ignored in this study due to the assumption of steady state conditions and to
simplify the analyses. No external loads (e.g. buildings, road traffic, etc.) have been considered.

Page 63 of 77
Riverbank Collapse Task 5 Report

Figure 41: An example of an SVSlope2D analysis (Transect No. 4212).

Figure 42: An example of the result of an SVSlope2D analysis (transect no. 5529).
Page 64 of 77
Riverbank Collapse Task 5 Report

Figure 43: Riverbanks that are vulnerable to slumping are identified (shown in yellow dots) after the analyses
(assuming a river level at 0 m AHD; blue contour lines show elevations that are at –10 m AHD or deeper with
one metre interval; red dot shows the approximate location of the occurrence of past slumping; dashed
yellow lines are areas of bedrock margins; and levees are represented by the thick green line).

In this case study, riverbanks with FoS below 1.05 are considered to be vulnerable and could collapse
in the long-term. Each result of the numerical modelling was studied in detail, and it is concluded that
shallow/planar type failures, not massive rotational or deep seat failures, can be expected to occur in
the long-term when the pool level is held constant at 0 m AHD. The primary cause of these potential
failures is the over-steepened riverbanks, which have slope angles much larger than the ' of the Silty
CLAY (i.e. 41°). The riverbanks with high slope angles, which appear to be stable in the short-term
due to the influence of total suction, may collapse in extreme rainfall events in the longer term. The
over-steepened riverbank can be found, in this case and predominantly near the deep holes and also
on the outside bend of meanders, are vulnerable to accelerated erosion. The placement of the Fill
layer, which acts as surcharge, is found to be the secondary factor, whilst the levees or embankments,
which are located >10 metres away from the waterline, are very unlikely to contribute to potential
failure.

Page 65 of 77
Riverbank Collapse Task 5 Report

Therefore, characterising the topographies of the riverbanks, such as the inclination angle of the
riverbank and the thickness of fill above pool level, using ArcGIS, can be used as indicators to aid with
high-risk areas identification. In the evaluation of landslide and riverbank slope failure in wide natural
slopes, it is typical to assume the slope surface is perfectly level. However, natural slopes are
generally undulating and irregular, an example is shown in Figure 42 and, as a consequence, the value
of the inclination can vary depending on the size of the grid that ArcGIS used to determined it. In
order to illustrate the effect of grid size, the inclination of the riverbanks near Thiele Reserve have
been mapped using the DEMs within ArcGIS and presented in Figures 44 and 45. Two different sizes
(or matrix) of grids: 1 × 1 in Figure 44 and 10 × 10 in Figure 45, are used in the DEM to evaluate the
inclination of the riverbank slope. By comparing Figures 44 and 45, it is observed that the size of the
adopted grid can greatly influence the evaluation of the inclination. For example, it can be seen that
flatter (10° to 30°) inclinations are obtained for vulnerable riverbanks when a 10 × 10 matrix of grids is
used (Figure 44), while a more detailed and variable (from a low value of 10° to more than 45°)
inclination map is obtained with a finer grid (e.g. 1 × 1 matrix of grids as shown Figure 44). Figure 44
clearly shows that the high-risk riverbanks are associated with high inclination values; therefore, the
inclination map determined by using a finer grid is accurate and more suitable to be used in the
identification of high-risk areas.

5.3 Summary
In summary, this section has shown that: (i) deep holes can be determined using DEM with ArcGIS; (ii)
deep holes can be found from Blanchetown to Wellington, but there a greater number of deep holes,
and they are longer and more widespread, downstream of Mannum; and (iii) the riverbanks near
Thiele Reserve, with slope angles greater than 40° are vulnerable and could collapse in the long-term,
but they are most likely to be shallow or planar failure types. The modelling has been undertaken
using the limit equilibrium method in SVSlope2D. To facilitate the stability analyses, a GIS framework
has been adopted and topographic information of the site has been obtained from a digital elevation
model (DEM). Finally, topographical factors, such as the inclination of the slope, can be used as a very
effective indicator for identifying riverbank instability near Thiele Reserve. The Task 6 report will
undertake further study and expand the study to include other locations.

Page 66 of 77
Riverbank Collapse Task 5 Report

45°

Figure 44: Inclination of the riverbanks near Thiele Reserve using a 1 × 1 matrix of grids.

Page 67 of 77
Riverbank Collapse Task 5 Report

45°

Figure 45: Inclination of the riverbanks near Thiele Reserve using a 10 × 10 matrix of grids.

Page 68 of 77
Riverbank Collapse Task 5 Report

6 Summary
This report presents the methodologies that have been developed and used by the authors to back-
analyse a number of historical collapses, was well as tools developed to assess the stability of the
riverbanks stabilities and identify the high-risk areas along the Lower River Murray. The geographic
information system (GIS) technique, implemented using ArcGIS, has greatly facilitated riverbank
instability research by providing a number of useful functions for spatial data management. The GIS-
based technique simplifies the process of identifying collapsed regions (locations, size and volume),
deepened riverbeds, over-steepened channels and the geometry and cross-sections of the riverbanks
of the Lower River Murray. Satellite images, aerial photographs and digital elevation models (DEMs)
have been provided by the South Australian Department of Environment, Water and Natural
Resources (DEWNR) for this research, and these high-resolution, remotely sensed data have been
processed using ArcGIS and studied extensively throughout this research.

This report has demonstrated that by integrating GIS-based methods and slope analysis tool (both
ArcGIS and SVSlope), riverbank collapses can be accurately and reliably predicted by using the
proposed frameworks. The effects of the rainfall and groundwater flow on the stability of the
riverbank has also been studied by integrating a groundwater and unsaturated soil modelling tool,
called SVFlux with the slope stability analysis software, SVSlope, both in 2D and 3D. The changes in
groundwater level and flow correspond to the changes of the river pool level. The moisture content
of the soil, which is affected by rainfall, evapotranspiration and ground water flow/level influences
the soils’ suction and shear strength, and ultimately riverbank stability. By validating the models, the
studies have demonstrated that the proposed framework provides reliable riverbank stability
predictions. The integration of GIS with high-resolution spatial data facilitates the process of
collapsed region identification, model geometry development and the calculation of the dimensions
of the collapsed regions.

The accuracy of the stability prediction hinges on the accuracy of soils parameters. Many factors can
influence the accuracy of the soil parameters determined from field and laboratory testing, such as
the field sampling technique, natural variability of the soil properties, the number of samples, sample
preparation and quality control in the laboratory. Therefore, back-analyses have been undertaken at
4 collapsed sites. The results of the back-analysed soil shear strengths at the 4 sites have shown great
consistency with those obtained from site investigations undertaken at locations adjacent to the
collapse sites.

Both 2D and 3D stability modelling yielded excellent predictions of the collapse when compared
against the recorded dates of collapse and dimensions of the failed region. However, due to project
time constraints, 2D analyses will be adopted for the Task 6 report, as the 2D modelling only requires
modest resources and provides reliable yet conservative estimates of FoS.

Page 69 of 77
Riverbank Collapse Task 5 Report

7 References

Abramson, LW, Lee, TS, Sharma, S and Boyce, GM, 2002, ‘Slope stability and stabilization methods’,
2nd edition, John Wiley & Sons, Inc., New York.
Acharya, G, De Smedt, F, and Long, N, 2006, ‘Assessing landslide hazard in GIS: a case study from
Rasuwa, Nepal’. Bulletin of Engineering Geology and the Environment, 65(1), 99–107.
Arnott, C, 1994, ‘Rates of stage fluctuation on the Murray River between Hume reservoir and Albury’.
Honours research thesis, University of Melbourne.
Aronoff, S, 1989, ‘Geographic Information System: A Management Perspective’. WDL Publications,
Ottawa, 279 pp.
Beal, A, Brown, R and Toole, JO, 2010, ‘Expert panel workshop – Riverbank collapse hazard program’.
(unpublished).
Berilgen, MM, 2007, ‘Investigation of stability of slopes under drawdown conditions’. Computers and
Geotechnics, 34(2), 81–91.
Bouwer, H, 1978, ‘Groundwater hydrology’. McGraw-Hill, New York.
Bozzano, F, Martino, S, Montagna, A and Prestininzi, A, 2012, ‘Back analysis of a rock landslide to infer
rheological parameters’. Engineering Geology, 131-132, 45–56.
Brand, EW, 1984, ‘Landslides in Southeast Asia: A state-of-the-art report’. Proceedings of 4th Int. Symp.
on Landslides, Toronto, 17–59.
Budhu, M and Gobin, R, 1995, ‘Seepage-induced slope failure on sandbars in grand canyon’. Journal of
Geotechnical Engineering, 121(8), 601–609.
Bureau of Meteorology, 2014, ‘Climate data online’, Commonwealth of Australia,
<http://www.bom.gov.au/jsp/ncc/cdio/weatherData/>.
Burns, SE and Mayne, PW, 1998, ‘Monotonic and dilatory pore pressure decay during piezocone tests
in clay’. Canadian Geotechnical Journal, 35, 1063–1073.
Burrough, PA, 1986, ‘Principles of geographical information systems for land resources assessment.’
Clarendon Press, Oxford, 194 pp.
Cai, MF, Xie, MW and Li, CL, 2007, ‘GIS-based 3D limit equilibrium analysis for design optimization of a
600 m high slope in an open pit mine’. Journal of University of Science and Technology Beijing,
14(1), 1–5.
Calabresi, G, Colleselli, F, Danese, D, Giani, GP, Mancuso, C and Montrasio, L, 2013, ‘A research study
of the hydraulic behaviour of the Po river embankments’. Canadian Geotechnical Journal, 50(9),
947–960.
Carrara, A, Cardinali, M, Detti, R, Guzzetti, F, Pasqui, V and Reichenbach P, 1991, ‘GIS techniques and
statistical models in evaluating landslide hazard’. Earth Surface Processes and Landforms, 16,
427–445.

Page 70 of 77
Riverbank Collapse Task 5 Report

Carrara, A, Cardinali, M, Guzzetti, F and Reichenbach, P, 1995, ‘GIS-based techniques for mapping
landslide hazard.’ In GIS in assessing natural hazards, Eds. A Carrara, A and F Guzzetti, Kluwer
Publications, Dordrecht, 135–176.
Casagli, N, Rinaldi, M, Gargini, A and Curini A, 1999, ‘Pore water pressure and streambank stability:
Results from a monitoring site on the Sieve River, Italy’. Earth Surface Processes and Landforms,
24(12), 1095–1114.
Coffey 2012, ‘Review of management options for four riverbank collapse high risk sites’. 07093AA-AD
Final Report, December 12, Adelaide.
Dahm, J and Edwards M, 1988, ‘Preliminary investigation of the influence of Matahina Power Station
on river bank stability along the Rangitaiki River’. Electricity Corporation of New Zealand Ltd.
Hamilton.
DFW, 2010. ‘Riverbank collapse hazard expert panel, submission: site reports & investigation’.
Department for Water, Government of South Australia. (unpublished).
DFW, 2011, ‘Lower River Murray riverbank collapse hazard site status report May 2011’.
Filz, GM, Brandon, TL, and Duncan, JM, 1992, ‘Back analysis of the Olmsted Landslide using
anisotropic strengths’. Transportation Research, Record 1343, TRB, National Research Council,
72–78.
Fredlund, DG, 2006, ‘Unsaturated soil mechanics in engineering practice’. Journal of Geotechnical &
Geoenvironmental Engineering, 132(3), 286–321.
Fredlund, DG and Xing, A. 1994, ‘Equations for the soil-water characteristic curve’. Canadian
Geotechnical Journal, 31(4), 521–532.
Fredlund, DG, Xing, A and Huang, S, 1994, ‘Predicting the permeability function for unsaturated soils
using the soil-water characteristic curve’. Canadian Geotechnical Journal, 31(4), 533–546.
Fredlund, DG, Xing, A, Fredlund, MD and Barbour, SL, 1996, ‘The relationship of the unsaturated soil
shear to the soil-water characteristic curve’. Canadian Geotechnical Journal, 33(3), 440–448.
Fredlund, DG, Xing, A and Huang, S, 1995, ‘Predicting the permeability function for unsaturated soils
using the soil-water characteristic curve’. Canadian Geotechnical Journal, 31, 533–546.
Gilbert, RB, Wright, SG and Liedtke, E, 1998, ‘Uncertainty in back analysis of slopes: Kettleman Hills
case history’. Journal of Geotechnical & Geoenvironmental Engineering, 124(12), 1167–1176.
Green, SJ, 1999, ‘Drawdown and river bank stability’. Masters research thesis, Dept. of Civil and
Environmental Engineering, University of Melbourne.
Harris, SJ, Orense, RP and Itoh, K, 2012, ‘Back analyses of rainfall-induced slope failure in Northland
Allochthon formation’. Landslides, 9(3), 349–356.
Hashimoto, A, Oguchi, T, Hayakawa, Y, Lin, Z, Saito, K and Wasklewicz, TA, 2008, ‘GIS analysis of
depositional slope change at alluvial-fan toes in Japan and the American Southwest’.
Geomorphology, 100(1-2), 120–130.
Higgins, DT, 1980, ‘Unsteady drawdown in 2D water-table aquifer’. Journal of the Irrigation and
Drainage Division, ASCE, 106(3), 237–251.

Page 71 of 77
Riverbank Collapse Task 5 Report

Hong, Y, Adler, R and Huffman, G, 2007a, ‘Use of satellite remote sensing data in the mapping of
global landslide susceptibility’. Natural Hazards, 43(2), 245–256.
Hooke, JM, 1979, ‘An analysis of the processes of river bank erosion’. Journal of Hydrology, 42(1-2),
39–62.
Hubble, T and De Carli, E, 2015, ‘Mechanisms and Processes of the Millennium Drought River Bank
Failures: Lower Murray River, South Australia’, Goyder Institute for Water Research Technical
Report Series No. 15/5, Adelaide, South Australia.
Institution of Engineers, Australia, 1987, ‘Australian rainfall and runoff: A guide to flood estimation’.
Vol. 1, Editor-in-chief D.H. Pilgrim, Revised Edition 1987 (Reprinted edition 1998), Barton, ACT.
Jaksa, MB, Kuo, YL, Hubble, T, De Carli, E and Liang, C, 2015, ‘Geotechnical investigations and data
acquisition’, Goyder Institute for Water Research Technical Report Series No. 15/x, Adelaide, South
Australia. (In press.)
Jia, GW, Zhan, TLT, Chen, YM and Fredlund, DG, 2009, ‘Performance of a large-scale slope model
subjected to rising and lowering water levels’. Engineering Geology, 106(1-2), 92–103.
Kamp, U, Growley, BJ, Khattak, GA and Owen, LA, 2008, ‘GIS-based landslide susceptibility mapping
for the 2005 Kashmir earthquake region’. Geomorphology, 101(4), 631–642.
Lane, PA, and Griffiths, DV, 2000, ‘Assessment of stability of slopes under drawdown conditions’.
Journal of Geotechnical & Geoenvironmental Engineering, 126(5), 443–450.
Lee, S, Choi, J and Min, K, 2004, ‘Probabilistic landslide hazard mapping using GIS and remote sensing
data at Boun, Korea’. International Journal of Remote Sensing, 25(11), 2037–2052.
Li, XP, Tang, HM and Chen, S, 2008, ‘A GIS-supported logistic regression model applied in regional
slope stability evaluation’. Landslides and Engineered Slopes: From the Past to the Future, vol. 1
and 2, 789–794.
Li, YQ, Feng, Y and Jiang, SH, 2011, ‘Stability analysis of soil slope during rapid drawdown of water
level’. Proceedings of 2nd Int. Conf. on Mechanics Automation and Control Engineering (MACE),
July 15-17, Hohhot, China, 3454–3457.
Liang C, Jaksa, MB and Ostendorf, B, 2012, ‘GIS-based back analysis of riverbank instability in the
lower River Murray’. Australian Geomechanics, 47(4), 59–65.
Liang, C, Jaksa, MB, Kuo, YL and Ostendorf, B, 2015a, ‘Back analysis of Lower River Murray riverbank
collapses’. Australian Geomechanics, 50(2), 29–41.

Liang, C, Jaksa, MB, Ostendorf, B and Kuo, YL, 2015b, ‘Influence of river level fluctuations and climate
on riverbank stability’. Computers and Geotechnics, 63(1), 83-98.
Liang, C, Jaksa, MB, Kuo, YL and Ostendorf, B, 2015c, ‘Identifying areas susceptible to high risk of
riverbank collapse along the Lower River Murray’. Computers and Geotechnics, 69(9), 236–246.

Lumb, P, 1975, ‘Slope failures in Hong Kong’. Quarterly Journal of Engineering Geology and
Hydrogeology, 8(1), 31–65.

Page 72 of 77
Riverbank Collapse Task 5 Report

Magliulo, P, Di Lisio, A, Russo, F and Zelano, A, 2008, ‘Geomorphology and landslide susceptibility
assessment using GIS and bivariate statistics: a case study in southern Italy’. Natural Hazards,
47(3), 411–435.
Marble, DF, 1990, ‘Geographic information systems: an overview – Introductory readings in
geographic information systems’. Eds. D. F. Marble and D. J. Peuquet, Taylor and Francis Ltd.,
New York, 9–17.
Mayo, D, 1982, ‘Permissible rates of river rise and fall below Dartmouth Regulating Dam.’ State
Electricity Commission of Victoria.
Merwade, VM, 2007, ‘An automated GIS procedure for delineating river and lake boundaries’.
Transactions in GIS, 11(2), 213–231.
Mukhlisin, M and Taha, M, 2012, ‘Numerical model of antecedent rainfall effect on slope stability at a
hillslope of weathered granitic soil formation’. Journal of the Geological Society of India, 79(5),
525–531.
Murray Darling Basin Authority, 2014, ‘Live river data – Murray river border to the sea’,
<www.mdba.gov.au/river-data/live-river-data/border-to-sea>.
National Water Commission, 2005, ‘Australian water resources 2005: A baseline assessment of water
resources for the National Water Initiative’, 115 pp., <http://www.water.gov.au/publications/
AWR2005_Level_2_Report_May07.pdf>.
Okagbue, C and Abam, T. 1985. ‘An analysis of stratigraphic control on the river bank failure.’
Engineering Geology, 22(3), 231–245.
Okui, Y, Tokunaga, A, Shinji, M and Mori, S, 1997, ‘New back analysis method of slope stability by
using field measurements’. Int. Journal of Rock Mechanics and Mining Sciences, 34(3–4), 234–249.
Osman, AM and Thorne, CR, 1988, ‘Riverbank stability analysis 1: Theory’. Journal of Hydraulic
Engineering, ASCE, 114(2), 134–150.
Ozdemir, A, 2009, ‘Landslide susceptibility mapping of vicinity of Yaka Landslide (Gelendost, Turkey)
using conditional probability approach in GIS’. Environmental Geology, 57(7), 1675–1686.
Pantha, BR, Yatabe, R and Bhandary, NP, 2008, ‘GIS-based landslide susceptibility zonation for
roadside slope repair and maintenance in the Himalayan region’. Episodes, 31(4), 384–391.
Pauls, GJ, Sauer, EK, Christiansen, EA and Widger, RA, 1999, ‘A transient analysis of slope stability
following drawdown after flooding of a highly plastic clay’. Canadian Geotechnical Journal, 36(6),
1151–1171.
Rahardjo, H, Leong, EC and Rezaur, RB, 2008, ‘Effect of antecedent rainfall on pore-water pressure
distribution characteristics in residual soil slopes under tropical rainfall’. Hydrological Processes,
22(4), 506–523.
Rahardjo, H, Li, XW, Toll, DG and Leong, EC, 2001, ‘The effect of antecedent rainfall on slope stability’.
Geotechnical & Geological Engineering, 19(3-4), 371–399.

Page 73 of 77
Riverbank Collapse Task 5 Report

Rahardjo, H, Ong, TH, Rezaur, RB and Leong, EC, 2007, ‘Factors controlling instability of homogeneous
soil slopes under rainfall’. Journal of Geotechnical & Geoenvironmental Engineering, 133(12),
1532–1543.
Rahimi, A, Rahardjo, H and Leong, EC, 2011, ‘Effect of antecedent rainfall patterns on rainfall-induced
slope failure’. Journal of Geotechnical and Geoenvironmental Engineering, 137(5), 483–491.
Ray, RL and Smedt, FD 2009, ‘Slope stability analysis on a regional scale using GIS: A case study from
Dhading, Nepal’. Environmental Geology, 57(7), 1603–1611.
Rinaldi, M and Casagli, N, 1999, ‘Stability of streambanks formed in partially saturated soils and
effects of negative pore water pressures: The Sieve River (Italy)’. Geomorphology, 26(4), 253–277.
Rinaldi, M, Casagli, N, Dapporto, S and Gargini, A, 2004, ‘Monitoring and modelling of pore water
pressure changes and riverbank stability during flow events’. Earth Surface Processes and
Landforms, 29(2), 237–254.
Schuster, RL and Krizek, RJ, 1978, ‘Landslides: Analysis and control’. Transportation Research Board,
National Research Council, Special Report No. 176.
SKM 2010a, ‘Study into river bank collapsing for Lower River Murray’, Geotechnical Investigation
Report.
SKM 2010b, ‘Study into river bank collapsing for Lower River Murray: Appendix A Riverfront Road –
Murray Bridge in-situ and laboratory test results’, Geotechnical Investigation Report.
SKM 2010c, ‘Study into river bank collapsing for Lower River Murray: Appendix B Caloote in-situ and
laboratory data’, Geotechnical Investigation Report.
SKM 2010d, ‘Study into river bank collapsing for Lower River Murray: Appendix C Woodlane Reserve
in-situ and laboratory data’, Geotechnical Investigation Report.
SKM 2010e, ‘Study into river bank collapsing for Lower River Murray: Appendix E East Front Road in-
situ and laboratory data’, Geotechnical Investigation Report.
SKM, 2011, ‘Research into processes, triggers and dynamics of riverbank collapse in the Lower River
Murray’.
SoilVision, 2009a, SVFlux, ver. 4.23, <http://www.soilvision.com>.
SoilVision, 2009b, SVSlope, ver. 4.23, <http://www.soilvision.com>.
Springer, FM, 1981, ‘Influence of the rapid drawdown events on river bank stability’. Masters research
thesis, University of Louisville.
Springer, FM Jr, Ullrich, CR and Hagerty, DJ, 1985, ‘Streambank stability’. Journal of the Geotechnical
Engineering Division, ASCE, 111(5), 624–640.
Springer, FM, 1981, ‘Influence of rapid drawdown events on river bank stability’, University of
Louisville.
Stark, TD and Eid, H, 1998, ‘Performance of three-dimensional slope stability methods in practice’.
Journal of Geotechnical and Geoenvironmental Engineering, 124(11), 1049–1060.

Page 74 of 77
Riverbank Collapse Task 5 Report

Stark, TD, 2003, ‘Three-dimensional slope stability methods in geotechnical practice.’ Proceedings of
51st Annual Geotechnical Engineering Conf., University of Minnesota, Minnesota Geotechnical
Engineering Group, Minneapolis, February, 41–74.
Sterrett, R and Edil, T, 1982, ‘Ground-water flow system and stability of a slope’. Ground Water, 20(1),
5–11.
Tajeddin, R, Ramsey, N, Pain, D, Mollison, D and Sandercock, P, 2010, ‘Influence of water level risk on
river bank collapse, Lower River Murray’. Sinclair Knight Merz, Australia.
Tang, WH, Stark, TD & Angulo, M, 1999, ‘Reliability in back analysis of slope failures’. Soils and
Foundations, 39(5), 73–80.
Thorne, CR and Osman, AM, 1988, ‘Riverbank stability analysis 2: Applications’. Journal of Hydraulic
Engineering, ASCE, 114(2), 151–172.
Thorne, CR, 1982, ‘Processes and mechanisms of river bank erosion.’ In Gravel-bed rivers, Eds. RD Hey,
JC Bathurst and CR Thorne, John Wiley & Sons, Chichester, 227–271.
Thorne, CR, Hey, RD, Newson, MD, 1997, ‘Bank erosion and instability’. In Applied fluvial
geomorphology for river engineering and management, Eds. CR Thorne, RD Hey and MD Newson,
John Wiley, New York, 137–172.
Torres Hernandez, G, 2011, 'Estimating the soil-water characteristic curve using grain size analysis and
plasticity index'. Masters thesis, Arizona State University.
Twidale, CR, 1964, ‘Erosion of an alluvial bank at Birdwood, South Australia’. Zeitschrift für
Geomorphologie NF, 8(2), 189–211.
Urgeles, R, Leynaud, D, Lastras, G, Canals, M and Mienert, J, 2006, ‘Back-analysis and failure
mechanisms of a large submarine slide on the ebro slope, NW Mediterranean’. Marine Geology,
226(3-4), 185–206.
Wang, H, Liu, G, Xu, W and Wang, G, 2005, ‘GIS-based landslide hazard assessment: An overview’.
Progress in Physical Geography, 29(4), 548–567.
Wang, L, Hwang, JH, Luo, Z, Juang, CH and Xiao, J, 2013, ‘Probabilistic back analysis of slope failure – a
case study in Taiwan’. Computers and Geotechnics, 51, 12–23.
WaterConnect, 2014, ‘Real time water data – Latest hourly observation', Government of South
Australia, <https://www.waterconnect.sa.gov.au/Systems/RTWD/SitePages/Home.aspx>.
WaterConnect, 2015, ‘Riverbank collapse historical incident website', Government of South Australia,
<https://www.waterconnect.sa.gov.au/Systems/RCHIW/Pages/Default.aspx>.
Wen, HJ, Li, X and Zhang, Jl, 2009, ‘An evaluation-management information system of high slope geo-
risk for mountainous city based on GIS’. Proceedings of 1st Int. Conf. on Information Science and
Engineering (ICISE 2009), Nanjing, China, December 26-28, 1975–1978.
Xie, M, Esaki, T and Zhou, G, 2004, ‘GIS-based probabilistic mapping of landslide hazard using a three-
dimensional deterministic model’. Natural Hazards, 33(2), 265–282.

Page 75 of 77
Riverbank Collapse Task 5 Report

Xie, MW, Esaki, T and Cai, MF, 2006, ‘GIS-based implementation of three-dimensional limit
equilibrium approach of slope stability’. Journal of Geotechnical and Geoenvironmental
Engineering, 132(5), 656–660.
Xie, MW, Zhou, GY and Tetsuro, E, 2003, ‘GIS component based 3D landslide hazard assessment
system: 3DSLOPEGIS’. Chinese Geographical Science, 13(1), 66–72.
Yan, ZL, Wang, JJ and Chai, HJ, 2010, ‘Influence of water level fluctuation on phreatic line in silty soil
model slope’. Engineering Geology, 113(1-4), 90–98.
Zaitchik, BF and Van Es, HM, 2003, ‘Applying a GIS slope-stability model to site-specific landslide
prevention in Honduras’. Journal of Soil and Water Conservation, 58(1), 45–53.
Zapata, CE, Houston, WN, Walsh, KD and Houston, SL, 2000, ‘Soil–water characteristic curve
variability’. Advances in Unsaturated Soil Mechanics, Geotechnical Special Publication No. 99, eds
CD Shackleford, SL Houston and N-Y Chang, Geo-Institute of ASCE, 84-124.
Zhang, JF, Li, ZG and Qi, T, 2005, ‘Mechanism analysis of landslide of a layered slope induced by
drawdown of water level’. Science in China Series E Engineering & Materials Science, 48(1), 136–
145.
Zhang, LL, Zhang, J, Zhang, LM and Tang, WH, 2011, ‘Stability analysis of rainfall-induced slope failure:
A review’. Proceedings of the Institution of Civil Engineers-Geotechnical Engineering, 164(5), 299–
316.
Zolfaghari, A and Heath, AC, 2008, ‘A GIS application for assessing landslide hazard over a large area’.
Computers and Geotechnics, 35(2), 278–285.

Page 76 of 77
Riverbank Collapse Task 5 Report

The Goyder Institute for Water Research is a partnership between the South Australian Government through the
Department of Environment, Water and Natural Resources, CSIRO, Flinders University, the University of Adelaide,
the University of South Australia and ICE WaRM (the International Centre of Excellence in Water Resources
Management).

Page 77 of 77

You might also like