Residential Footing Design On Expansive Soils - Sun Et Al (EJGE 2017 Vol 22)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Residential Footing Design on

Expansive Soils: A Review of


Australian Practice
Xi Sun
Postgraduate student, School of Engineering, RMIT University, Melbourne
3001, Australia
*
Corresponding Author, e-mail: xi.sun@rmit.edu.au

Jie Li
Associate Professor, School of Engineering, RMIT University, Melbourne
3001, Australia jie.li@rmit.edu.au

Gang Ren
Senior Lecturer, School of Engineering, RMIT University, Melbourne 3001,
Australia gang.ren@rmit.edu.au

ABSTRACT
The vast majority of Australian homes were made of one- or two-storey brick veneer, consisting of
an internal timber structure covered with a single layer of brick on the outside. Such lightweight
structures are vulnerable to ground movements as a result of cyclic soil shrinkage and swelling
arising from soil moisture (suction) variations. The Australian Residential Slabs and Footings
Standard AS2870 recommends the use of Thornthwaite Moisture Index (TMI) to quantify the effect
of climate changes on the residential footing design. The TMI based isopleth map for the State of
Victoria is endorsed by the standard, which assists geotechnical engineers to estimate the depth of
design soil suction change (Hs) needed for the calculation of the characteristic ground movement (ys),
which forms the basis for site classification for potential ground movements. This paper presents a
review of the current footing design and construction in Australia, including site classification
process and two commonly used footing types. The related design parameters in the Australian
Standards AS2870 are also introduced and discussed, which include the characteristic design surface
movement (ys), Thornthwaite Moisture Index (TMI), the depth of design suction change (Hs) and the
instability index (Ipt). A worked example for site classification is presented as well.

- 3939 -
Vol. 22 [2017], Bund. 10 3940

KEYWORDS: Expansive Soil, Australian Standard, Residential Footing, Thornthwaite

Moisture Index, Ground Movement.

INTRODUCTION
Expansive soils are clay soil or reactive soil that undergoes appreciable volume change as a
result of changes in soil moisture content (suction) due to natural effects such as normal seasonal
change or tree root activity, or man-made effects such as garden watering, leaking underground
water services, or a deficient storm water drainage system (Li et al., 2013). Seasonal wetting and
drying of expansive soils in response to climate cycles lead to shrinkage and swelling of soils that
can have significant impacts on the performance of residential buildings.

Damage to lightweight buildings caused by expansive soil movements has been widely reported
in many countries such as Australia, China, India, Israel, South Africa, the United Kingdom and the
United States of America (Li and Cameron, 2002; Li et al., 2014). Expansive soils have long been
of great concern to design engineers and practitioners in Australia. It was estimated that
approximately 20% of the surface area in Australia (Figure 1) is covered by expansive clays
(Richards et al., 1983; Holland, 1981). In fact, the expansive soil problems are present in most
capital cities and regional centers of Australia (Delaney et al., 2005).

The Australian Standard AS2870 was published of the need to regulate different practices
across States and to lessen the burden on the community of the costs of damage to lightweight
structures. This standard introduced a soil suction based method for estimating the free surface
movement (ys) in expansive soil, which has been widely employed by geotechnical engineers for
routine footing design with success over the past two decades. The climate indicator, Thornthwaite
Moisture Index (TMI), has been used by engineers to understand general differences in climate
across the nation and to estimate the depth of design soil suction change (Hs). AS2870 provides a
climate zoning map for the State of Victoria based on TMI, and assigns the depth of seasonal soil
moisture change Hs to each zone based on the early pioneering work by Aitchison and Richards
(1965) and Smith (1993) for the determination of characteristic ground movements ys, which forms
the basis for site classification needed for construction of footings. In this study, the ys calculation
procedure and site classification process are described.
Vol. 22 [2017], Bund. 10 3941

Figure 1: Distribution of moderately to highly expansive soils in Australia


(Richards et al., 1984)

SITE CLASSIFICATION
Research into expansive soil behavior has been carried out in Australia since the late 1950’s, the
outcome of the research efforts led to the establishment of a national standard for residential slabs
and footings design in Australia. The Australian Residential Slabs and Footings Standard AS2870
was first published in 1986, then revised and published in two parts in 1988 and 1990, followed by
two more complete editions in 1996 and 2011. Common to the all five versions of the standard, sites
are classified according to soil profile and regional climate influence on soil moisture state (Li et al.,
2014).

The footing design procedure provided in AS2870 is based on the site classification. AS2870
classifies sites based on the expected ground surface movement (i.e. the characteristic design
surface movement), ys, which is based on design soil suction change profiles for different climatic
regions of Australia (Li and Cameron, 2002). For sites where ground movement is predominately
due to soil reactivity, and where abnormal moisture conditions are not expected, the standard
provides for sites to be classified into one of five classes from slightly reactive to extremely reactive
in accordance with Table 1. Localities having Hs of equal to or greater than 3 m are categorized as
deep-seated moisture variation sites and further classified as M-D, H1-D, H2-D and E-D. Deemed
to comply footing designs were provided in AS2870 for most wall constructions. Case by case
Vol. 22 [2017], Bund. 10 3942

design are required for extremely reactive (E) sites. The standard also outlined an alternative
classification method of a site based on reasonable interpretation of the footing performance of
existing buildings on a similar soil of at least 10 years old.

Table 1: Site classification by ys (AS2870, 2011)


Characteristic surface Site classification Description
movement (ys) mm
0 < ys ≤ 20 S Slightly reactive
20 < ys ≤ 40 M Moderately reactive
40 < ys ≤ 60 H1 Highly reactive
60 < ys ≤ 75 H2 Very Highly reactive
ys > 75 E Extremely reactive

ESTIMATION OF THE CHARACTERISTIC GROUND

MOVEMENT
The characteristic ground movements ys is the moisture variation induced vertical movement of
the ground surface in a reactive soil site. It has less than 5% chance of exceeding ys in the life of the
structure (e.g. 50 years). The expected free surface movement ys is determined by estimating the
movement of each soil layer within the depth of deign suction change Hs and summing the
movement for all layers, as follows:

=ys
1 N
100 n =1
(
∑ I pt ∆uh ) n
(1)

where, Ipt is the instability index, indicating the degree of soil reactivity. It is defined as the axial
strain of the soil under a unit logarithm suction change, taking into account the expected values of
the applied stress, degree of lateral restraint and the soil suction range. It can be determined by
multiplying a shrinkage index Ips, by an empirical correction factor α (varying between 1 and 2) to
allow for full or partial lateral confinement, depending on the depth within the soil layer and the
expected depth of seasonal shrinkage cracking. For the majority of expansive soils, Ips values
generally fall in the range of 3 to 8 % strain per log(kPa). A shrinkage index of 4%/log(kPa) would
be regarded as a highly expansive soil, 6%/log(kPa) very highly expansive and 8%/log(kPa), an
extremely expansive soil (Li et al., 2014). ∆u is the soil suction change averaged over the thickness
of the layer under consideration, which is assumed to linearly decrease with depth. The maximum
suction change at the ground surface, ∆us is given in AS2870 for various cities and regions in
Australia. h is the thickness of layer and N is the number of layers within the depth of Hs. Values of
Vol. 22 [2017], Bund. 10 3943

Hs for various populated cities in Australia are provided in AS2870 and for other regions, the site
classifier has to have recourse to a Thornthwaite Moisture Index (TMI) map and use TMI-Hs
correlation (Table 2) to obtain the depth of design soil suction change Hs.

SHRINKAGE INDEX DETERMINATION FROM

LABORATORY TEST
Three laboratory test methods, namely the shrink-swell test (AS1289.7.1.1, 2003), the loaded
shrinkage test (AS1289.7.1.2, 1998b) and the core shrinkage test (AS1289.7.1.3, 1998c) can be used
to determine the shrinkage index for expansive soils. The core shrinkage test requires an
undisturbed cylindrical core sample to shrink under air-drying condition. Changes in the length and
the mass of the samples are measured periodically. The initial soil suction and moisture content of
the sample are usually determined by testing the sample trimmings. To convert moisture content to
suction, a further set of data is required to define the soil moisture characteristic (i.e. the slope of the
moisture content versus suction curve). This data may be obtained by drying small sub-samples in
vacuum desiccators containing saturated salts with the known suction. Considering the difficulty of
measuring changes in sample length due to shrinkage cracking and the difficulty of determining the
moisture characteristic accurately, the core shrinkage test is not a preferred method (Li and Zhou,
2013). An alternative to the core shrinkage test is the loaded shrinkage test, which uses a much
smaller cylindrical sample (approx. 38 mm in diameter and 25 mm in height), held under a vertical
pressure of either 25 kPa or the estimated overburden pressure (through a spring-load ram). The
spring-loaded sample is placed over copper sulphate solution (approximately 4.5 log10 (kPa)) within
a vacuum desiccator. This test requires considerable time (8 weeks or more) and is not suitable if
the soil is initially too dry as little if any shrinkage will occur. Of these test methods, only the
shrink-swell test deals with both swelling and shrinkage, and as such is more popular. The shrink-
swell test has two distinct advantages when compared to the other two methods: (a) both swell and
shrinkage strains are considered so that the sample may be either very wet or very dry; (b) there is
no need to measure soil suction values (Li et al., 2016).

The shrink-swell test consists of a core shrinkage test and a swelling test, which requires two
"companion" soil samples having the identical initial moisture content. The core shrinkage test
requires a laterally unrestrained and unloaded cylindrical core sample of a diameter of 45-50 mm to
be cut and trimmed to a length within the range of 1.5 to 2 diameters. The soil sample is first
allowed to be air dried on a smooth surface and then oven dried to a constant mass at 105°C
degrees. The changes in sample length, diameter and mass are monitored with time. In the swell
test, a sample of a 50 mm diameter is cut using a rigid steel ring of 20 mm height and 45 mm in
diameter and trimmed carefully to ensure both ends are flat. The sample is placed in a consolidation
cell with two porous stone plates mounted at the top and bottom. Initially a seating pressure of 5 kPa
Vol. 22 [2017], Bund. 10 3944

is applied for about 10 minutes and the displacement transducer was zeroed under this seating load
to allow for a small amount of initial settlement of specimen. The vertical pressure is then increased
to a value equal to the overburden pressure or 25 kPa for a maximum period of 30 minutes
(AS1289.7.1.1, 2003) following by adding stilled water to the soil to make it swell.

Once both shrinkage strain and swelling strain are determined, the shrink-swell index can be
calculated using the following expression (AS1289.7.1.1, 2003)

ε sw
ε sh +
I ss = 2.0 (2)
1.8

It can be seen from Equation 2, the swelling strain, εsw is divided by an empirical factor of 2 to
correct for lateral restraint, although different values of correction factor are proposed by
researchers. Finding by Fityus (1996) through the study of clay soils from Newcastle, New South
Wales, suggesting that the factor of 2 is appropriate. In order to calculate the swell-shrink index, the
total strain is divided by an empirical parameter of 1.8, which is the assumed range of soil suction
change.

THE THORNTHWAITE MOISTURE INDEX


Thornthwaite formulated his climatic classification system in terms of general descriptive field
observations in 1931 (Thornthwaite, 1931). In 1948, he pioneered a climate parameter, the
Thornthwaite Moisture Index, which was derived from P-E Index (i.e. precipitation minus
evaporation) to supersede the means of climate classification. TMI was quickly accepted by the
scientific community and has been employed across a diverse range of scientific research areas in a
vast number of applications. Initially, the TMI was employed to map the climate zones of the
United States of America (Thornthwaite, 1948). Following this concept the TMI has been used to
map the extent and characteristics of ecological communities. Aitchison and Richards (1965)
adopted the TMI to estimate potential soil suctions beneath covered areas. TMI is used as an
indicator of suitable climates for various crops types in the agricultural field. Perera et al. (2004)
developed a new model to predict suctions beneath pavements using the TMI. The index has also
seen application in predicting water catchment capacities and forest growth. In Australia, TMI has
been used by engineers to understand general differences in climate across the nation and estimating
the depth of design soil suction changes Hs which is needed for the determination of characteristic
ground movement, ys.
Vol. 22 [2017], Bund. 10 3945

TMI is mainly a function of rainfall and potential evapotranspiration (PE). The calculation of
TMI is on a yearly basis and expressed in a number which is computed for a statistically significant
number of years to allow a true climatic condition to be obtained for a particular region. A negative
TMI indicates an arid climate with a deficit of precipitation relative to PE and generally low
moisture in the soil. A positive TMI indicates a humid climate with a surplus of precipitation and
generally high soil moisture. Zero TMI index means that, over the long term, under average
conditions, the inflow from precipitation just equals the loss of soil moisture through
evapotranspiration that would occur if the notional reference vegetative cover was actually present.

TMI computation equation originally proposed by C. W. Thornthwaite (1948) is expressed as


follows:

TMI= I h − 0.6 I a (3)

where Ih and Ia are indices of humidity and aridity respectively, and are given by:

100 R 100 D
=Ih = Ia (4)
PE PE

where R represents the moisture surplus or runoff (mm) (i.e. the amount of rainfall that cannot
infiltrate a wet site); D represents the moisture deficit (mm) (i.e. the quantity of water that cannot be
evapotranspired from a dry site because it is not available) and PE is the adjusted potential
evapotranspiration (mm) and represents the water need.

The coefficient of water surplus and deficit is based on the assumption that, 6 inches (15.24 cm)
moisture surplus in one season will compensate 10 inches (25.4 cm) deficiency in another season,
and the reason, as explained by Thornthwaite (1948), is simply because it is easier for water to enter
the soil profile than it is for water to be extracted.

The original TMI equation (Equation 3) has been widely adopted for the TMI calculation by
many researchers (e.g. McKeen and Johnson, 1990; Fox, 2000 and 2002; Chan and Mostyn, 2008
and 2009; Jewell and Mitchell, 2009; Lopes and Osman, 2010 and Mitchell, 2012 and 2013). It was
employed by Aitchison and Richards (1965) to develop the first TMI isopleth map for Australia.
Based on the work of Aitchison and Richards (1965), Smith (1993) created the TMI-based climate
maps for the State of Victoria and its capital city Melbourne, which have been incorporated with
some committee-sanctioned changes in AS2870 (1996, 2011) as Figures D1 and D2.
Vol. 22 [2017], Bund. 10 3946

It can be seen from Equation 4 that three parameters are needed to be determined to derive TMI,
which are monthly moisture surplus (R), monthly moisture deficit (D) and monthly adjusted
potential evapotranspiration (PE) respectively. The first two parameters can be computed using the
water balance approach and the detailed calculation procedure is given by McKeen and Johnson
(1990). The initial (S0) and maximum (Smax) water storage value are usually taken to be respectively
0mm and 100mm in order to carry out the water balance analysis.

There are two different methods for the determination of TMI, namely the ‘year-by-year’
method and the ‘average year’ method. Most researchers (e.g. Fox, 2000 and 2002; Chan and
Mostyn, 2008 and Lopes and Osman, 2010) have adopted the first method to derive TMI. Sun et al.
(2017) used different values of S0 to examine the effect of the initial water storage S0 on the use of
two different methods on the TMI calculations, and found that the ‘average year’ method was very
susceptible to the assumed S0 and thus this method should be avoided while the ‘year-by-year’
method was not sensitive to the values of S0.

POTENTIAL EVAPOTRANSPIRATION (PE)


Potential evapotranspiration (PE) is the soil evaporation and crop transpiration under sufficient
water supply. It is the measurement of the ability of the atmosphere to transport water from the
surface through the evaporation and transpiration processes. This term was first introduced by
Thornthwaite in 1948 to define the evapotranspiration that would occur when there was an adequate
supply of soil moisture at all times (Chow, 1964).

The Thornthwaite method (Thornthwaite, 1948) for determining the adjusted potential
evapotranspiration (PEi) for the month, i, is expressed as:

DN 
PEi = ei  i i  (5)
 30 

where Di is the day length correction factor for the month i; Ni is the number of days in the month i;
and ei is the non-adjusted potential evapotranspiration (cm) for the month i given as:

a
 10t 
ei = 1.6  i  (6)
H
 y 
Vol. 22 [2017], Bund. 10 3947

where ti is the mean monthly temperature in °C and calculated as the average of tmax and tmin. The
heat index for each month is determined as follows:

hi = ( 0.2ti )
1.514
(7)

The annual heat index, Hy is simply determined by summing the 12 monthly heat index values.
The power term a in Equation 6, is given as:

a= 6.75 × 10−7 H y3 − 7.71 × 10−5 H y2 + 0.017921H y + 0.49239 and 0 < a < 4.25 (8)

PE plays a critical role in TMI calculation since TMI indices can vary greatly if different PE
estimation methods are employed. There are approximately 50 models available for estimation of
evapotranspiration (Grismer et al., 2002), ranging from very simple equations which require only
one or two meteorological parameters to more sophisticated and accurate models. Of these models,
A relatively simple temperature-based PE calculation method (Thornthwaite, 1948), which was
developed based upon global climate pattern distribution and the concept of plant physiology
relating to moisture availability (Jewell and Mitchell, 2009), has been used by a number of
researchers for the purpose of TMI calculation, these include Fityus et al. (1998); Fox (2000, 2002);
Chan and Mostyn (2008, 2009); Jewell and Mitchell (2009); Mitchell (2012, 2013); Er and Rifat
(2014) ; Li and Sun (2015) and Karunarathne et al. (2016).

THE DEPTH OF DESIGN SOIL SUCTION CHANGE HS


Soil suction profile (Figure 2) was a representation of a state of physical balance between
various processes operating to add or to subtract water at any part of the profile (Aitchison and
Woodburn, 1969). The design suction change profiles were simple approximations to what had been
found from site investigations, and can be obtained from suction measurements at various depths
during wet and dry periods. The simplified reverse triangular suction profile has been adopted in
AS2870 and it assumes the reduction of suction change is linear with depth to the maximum depth,
Hs. However, the observed wet and dry suction values usually decrease rapidly with depth (Fityus et
al., 1998).
Vol. 22 [2017], Bund. 10 3948

Figure 2: Typical soil suction profile (Coffey, 1985)


The soil profile is considered in conjunction with simplified design suction changes to produce
estimates of ground movement for footing design. The depth of design soil suction change Hs, is the
depth below which has no significant seasonal suction changes and hence no soil volume changes
occur in response to climate. Hs data obtained from field measurement are very rare and regionally
specific, which means its application range is very limit. Ideally, the value of Hs is determined by
long-term site monitoring of soil profile behavior which involves extremely dry and wet conditions.
However, collecting data of ground movement, soil suction and moisture changes may take years
even decades, which is not only time-consuming but also a tough and challenging work to perform.

Fortunately the early work by Aithison and Richards (1965) have found that Hs correlates well
with the climate index, TMI. In AS2870, the approach which was originally proposed by Smith
(1993) is adopted to estimate Hs (Fityus and Olivier, 2008). AS2870 (1996) recommended that at
least 20 continuous years of climate data are needed for the estimate of Hs while the current version
(AS2870, 2011) suggests to use 25 continuous years of climate data records as the datum for TMI
calculation.
Vol. 22 [2017], Bund. 10 3949

Table 2: Correlation between TMI, Hs and climatic zone (AS2870, 1996 and 2011)
Climatic TMI
Description Hs (m)
Zone AS2870(1996) AS2870(2011)
1 Alpine / Wet coastal > +40 > +10 1.5
2 Wet temperate +10 to +40 -5 to +10 1.8
3 Temperate -5 to +10 -15 to -5 2.3
4 Dry temperate -25 to -5 -25 to -15 3.0
5 Semi-arid < -25 -40 to -25 4.0
6 Arid - < -40 > 4.0

The correlation between TMI and Hs outlined in AS2870 is shown in Table 2 and Figure 3. It
can be seen that the standard changed the TMI range of each zone from 1996 to 2011 and an
additional climate Zone 6 was added to AS2870 (2011) edition, indicating the arid climate zone
where its corresponding Hs could exceed 4 m (Figure 4) due to the severe aridity of the climate.

Comparing TMI indices for two versions of the standard, it is noted that the values of each zone
are reduced to accommodate the trend of drying climate in Victoria. The threshold of climate Zone
1 has been changed from TMI >+40 (AS2870, 1996) to TMI >+10 (AS2870, 2011). This
amendment is based on the field observation that ground movement is very unlikely to arise in a wet
and humid climate (Lopes and Osman, 2010). Climate Zone 3 with TMI ranging from -5 to +10 in
the 1996 edition has been re-classified as Zone 2 (refer to Table 2) in the AS2870 (2011). In
addition, TMI of -15 has been introduced as a new threshold to reduce the TMI range from -25
≤TMI ≤-5 (Zone 4, 1996 edition) to -15 ≤TMI ≤ -5 for climate Zone 3 (2011 edition) and -25 ≤TMI
≤ -15 for climate Zone 4 (2011 edition) respectively. It is worth mentioning that in the 1996 edition,
Hs range of 1.5m - 2.3m was prescribed for Melbourne whereas in the current edition, Hs values are
increased slightly to a range of 1.8m - 2.3 m.
Vol. 22 [2017], Bund. 10 3950

5.0

4.0
Hs(m)

3.0

2.0

1.0

0.0
-60 -40 -20 0 20 40
Thornthwaite Moisture Index

Figure 3: Relationship between TMI and Hs by AS2870 (2011)

Surface soil suction change ∆us =1.2 log10(kPa) for all climate
Depth of design suction change Hs (m)

Zone 1
Zone 2
1.5 Zone 3
1.8 Zone 4
2.3 Zone 5
Zone 6
3.0

4.0
>4.0

Figure 4: Typical soil suction change profiles for climate zones (AS2870, 2011)
Vol. 22 [2017], Bund. 10 3951

RESIDENTIAL FOOTING DESIGN UNDER AS2870


There are two common types of footing construction in Australia, namely waffle raft slab and
stiffened raft slab (Figure 5). The conventional stiffened raft concrete slab consists of steel
reinforced concrete beams embedded in the soil which are cast integrally with a minimum 100mm
thick steel reinforced floor slab. This type of slab can be used for most classes of sites. Compare to
stiffened raft, waffle slab does not require any excavations and it made with polystyrene pods
placed in a grid formation, typically 1m by 1m, and separated by 110mm wide internal steel
reinforced concrete beams to support a 85mm steel reinforced floor slab. Waffle slabs can be used
for sites with slightly to moderately reactive clay but not recommend for highly reactive sites since
they are constructed on a level ground platform and are more susceptible to soil moisture variation
induced ground surface movement.

(a) Stiffened Raft Slab (b) Waffle Raft Slab

Figure 5: The two most commonly used footing types in Australia

The design of a slab footing can proceed with a given house construction type and the
calculated characteristic surface movement ys, using conventional two-dimensional design
principles. The irregularly shaped house plan is divided into a number of overlapping rectangles.
The design is focused on two basic mound shapes, i.e. short-term “edge heave” and long-term
“center heave” as illustrated in Figure 6. The worst case governs the design of the whole house. For
example, a stiffened raft slab is designed for a house and one of the rectangles requires the deepest
beams for center heave and thus all beams over the entire raft slab will comply with this design.
Design has to meet three criteria, including adequate footing stiffness, flexural strength and
sufficient section ductility.
Vol. 22 [2017], Bund. 10 3952

(a) Preliminary stage - edge heave

(b) Desiccation cycle - center heave

Figure 6: Typical slab deformation during periodic drying and wetting

It was recognized that post-construction site management is essential. For home owners had
their footing designed and constructed in accordance with AS2870, good site maintenance are
recommended to make sure abnormal soil moisture changes around their property over the life of
the house (i.e. 50 years) are avoided since such moisture variations can cause severe soil swell and
the subsequent excessive ground movements. Examples of avoidable moisture changes including
overwatering of gardens, unnoticed drainage pipe leaks and inadequate site drainage. Trees or other
substantial vegetation which are planted close to a house may affect the soil suction profile and thus
lead to increases in the ys. Assessment of the impact of vegetation on footing design is beyond the
scope of this paper.

ANALYSE TMI INDICES FOR DENSELY POPULATED

CITIES IN VICTORIA
TMI indices have been calculated and analyzed since 1954 to assess the changes in climate for
Melbourne and Geelong, the top two densely populated cities in Victoria. Fully recorded historical
long-term precipitation and temperature data from Melbourne Regional Office (No. 086071) and
Geelong SEC (No. 087025) weather stations were purchased from the Bureau of Meteorology
(BOM). The potential evapotranspiration (PE) values were calculated using the Thornthwaite
evapotranspiration equation (Equation 5). The original Thornthwaite equation (i.e. Equation 3) in
conjunction with the ‘year-by-year’ analysis method were employed for TMI computation. The
initial (S0) and maximum water storage (Smax) values were taken to be 0mm and 100 mm
respectively.
Vol. 22 [2017], Bund. 10 3953

Annual precipitation and TMI variation of Melbourne City for two consecutive 30-year periods
(i.e. 1954-1983 and 1984-2013) depicted in Figures 7-8 suggest the patterns of TMI and annual
precipitation changes are fairly similar although there are significant variations about the mean
values. The average TMI for Melbourne City decreased from -5 (1954-1983) to -12 (1984-2013)
while the mean precipitation reduced from 66 cm to 62 cm for the same period, which indicates
increasing drying of Melbourne’s climate since 1954. During the period of 1954-1983, the highest
annual total precipitation of 87 cm was recorded in 1978 and the rainfall amounts have declined in
an increasingly pronounced manner for the following year in 1979. During the second 30-year
period, the highest and lowest annual total precipitation was 84 cm and 36 cm, occurred in 1993 and
1997 respectively. It is interesting to note that the substantial positive TMI value of 11 (1989) did
not correspond with the highest annual total rainfall (84 mm in 1993). This can be attributed to the
effect of relatively lower rates of potential evapotranspiration in 1989. It can be seen from Figure
8(a) that there were four consecutive drought years from 2006 to 2009 with TMI (varying from -26
to -28) significantly lower than the mean TMI of -12. In December 2011, the Housing Industry
Association (HIA) estimated that more than 1000 new houses in the western suburbs of Melbourne
were damaged due to soil heave (The Age, 2011). These new houses have all been built in drier
weather between 2003 and 2011 and have been subjected to larger ground differential movements
induced by abnormal moisture changes after the construction of gardens/lawns and watering system
around buildings and the breaking of long drought in 2011.

(a) Variation in annual rainfall and TMI

Figure 7: Continues on the next page


Vol. 22 [2017], Bund. 10 3954

(b) Correlation between annual rainfall and TMI


Figure 7: Melbourne, 1954-1983

Annual precipitation and the corresponding TMI for Melbourne are plotted in Figures 7(b) and
8(b), which show a demonstrable linear relationship although variations about the trend line are
observed. The coefficient of determination (R2) is a statistical measure of how close the data are to
the fitted regression line, it has a range from 0 (no correlation) to 1 (perfect correlation). R2 of 0.9
was obtained for 1984-2013 period, indicating a strong positive correlation between annual TMI
and precipitation while a slightly smaller R2 was obtained for the first 30-year period.
Vol. 22 [2017], Bund. 10 3955

(a) Variation in annual rainfall and TMI

(b) Correlation between annual rainfall and TMI

Figure 8: Melbourne, 1984-2013

The gradual drying in the climate of the second largest city of Victoria, the City of Greater
Geelong, has also occurred from 1954 to 2013. The annual precipitation and TMI variation trends of
Geelong for two 30-year intervals plotted in Figures 9-10, suggesting that TMI values are highly
depended on precipitation and these indices have almost same variation as annual precipitation
although both parameters varied significantly about the mean values. Similar to Melbourne, a
marked reduction in mean annual precipitation and TMI has occurred in the past 30 years. As can be
seen from Figure 10(a), the TMI indices for Geelong were well below zero in the last three decades,
Vol. 22 [2017], Bund. 10 3956

which clearly indicates that soils in Geelong have progressively become drier for the period of
1984-2013. It is worth mentioning that the total amount of rainfall in three consecutive years from
1984 to 1986 was exactly the same but their corresponding TMI values varied from -16 to -20. This
can be attributed to the effect of temporal distribution of precipitation for years under consideration.
A good correlation between annual TMI and precipitation for all study periods is demonstrated in
Figures 9(b) and 10(b).

(a) Variation in annual rainfall and TMI

(b) Correlation between annual rainfall and TMI


Figure 9: Geelong, 1954-1983
Vol. 22 [2017], Bund. 10 3957

(a) Variation in annual rainfall and TMI (b) Correlation between annual rainfall and
TMI
Figure 10: Geelong, 1984-2013

A WORKED EXAMPLE FOR SITE CLASSIFICATION


A worked example is given below to show how to classify a site using TMI, Hs, Ips and ys. The
site is located in Braybrook, a western suburb of Melbourne. This site contains no apparent
geological anomalies.

Step 1: Calculation of TMI for the study period

Monthly precipitation and mean monthly maximum and minimum temperature data extracted
from Melbourne Regional Office (No. 086071) were used for TMI calculation for the study length
of 30 years (i.e. 1984-2013). S0 of 0 mm and Smax of 100 mm were adopted for initial and maximum
water storage respectively. Mean annual TMI (1984-2013) of -12 was computed for Braybrook
using estimated yearly values of R, D and PE from Equations 3, 4 and 5 in conjunction with the
‘year-by-year’ approach. An example of PE and TMI calculation for 2010 is given in Table 3 and 4.
Vol. 22 [2017], Bund. 10 3958

Table 3: PE calculation for Braybrook in 2010


Station Name: Melbourne CBD Station No.: 086071 Latitude: 37.81oS
No. of Day
days in length Precipitation Mean
Max.
Mean Mean Heat
Min. Avg. Index Unadjusted Adjusted
Month month Corr. (cm) Temp. Temp. Temp. (h ) a PE (cm) PE (cm)
(Ni) factor (P) (℃) (℃) (℃) i (ei)
(Di)
Jan 31 1.2525 2.56 27.5 15.9 21.70 9.23 9.60 12.42
Feb 28 1.0512 6.28 28.2 18.2 23.20 10.21 10.75 10.55
Mar 31 1.0656 8.48 26.2 16.3 21.25 8.94 9.26 10.20
Apr 30 0.9344 2.14 22.9 14.5 18.70 7.37 7.46 6.97
May 31 0.8731 2.50 18.2 9.8 14.00 4.75 4.57 4.12
Jun 30 0.7975 6.34 14.6 7.5 11.05 3.32 3.06 2.44
1.69
Jul 31 0.8531 3.08 14.6 7.3 10.95 3.28 3.01 2.65
Aug 31 0.9288 6.82 14.7 7.6 11.15 3.37 3.10 2.98
Sep 30 1.0000 5.30 16.5 9.1 12.80 4.15 3.92 3.92
Oct 31 1.1412 14.48 20.7 10.9 15.80 5.71 5.61 6.61
Nov 30 1.1869 11.52 23.2 13.4 18.30 7.13 7.19 8.53
Dec 31 1.2725 8.52 24.8 14.5 19.65 7.94 8.11 10.67
Sums 78.02 75.40 82.06

Table 4: Determination of TMI for Braybrook in 2010


Moisture
Moisture Moisture Moisture
Month Deficit TMI
Change(cm) Storage (cm) Surplus (cm)
(cm)
Jan -9.86 0.00 9.86 0.00
Feb -4.27 0.00 4.27 0.00
Mar -1.72 0.00 1.72 0.00
Apr -4.83 0.00 4.83 0.00
May -1.62 0.00 1.62 0.00
Jun 3.90 3.90 0.00 0.00
Jul 4.33 4.33 0.00 0.00 -4
Aug 8.17 8.17 0.00 0.00
Sep 9.54 9.54 0.00 0.00
Oct 17.41 10.00 0.00 7.41
Nov 12.99 10.00 0.00 2.99
Dec 7.85 7.85 0.00 0.00
Sums 22.29 10.40
Vol. 22 [2017], Bund. 10 3959

Step 2: Establish soil suction change profile

The Hs of 2.15m is obtained by using TMI-Hs correlation given in Table 2. To determine ys, the
shrink-swell indices, Iss, is required for each unique soil layer throughout Hs. Soil suction change
profile (Figure 12) is delineated based on the shrink-swell indices Iss obtained from the shrink-swell
test and soil profile (Figure 11) for Braybrook site.

0.0 ∆us =1.2log10(kPa)


0.0
Clay (CH), Grayish Brown, Stiff, root
fibres present. I =4.53% Iss = 4.53%
ss
Depth (m)

0.9 0.9

Hs (m)
Clay (CH), Brown to dark grey, Stiff, Iss = 5.52%
Slight calcareous. I =5.52% 1.61
ss
Limit of
cracking
2.0 2.0
Clay (CH), Light grey, Very stiff, Very
2.15
slight calcareous. I =4.90% Iss = 4.90%
ss
2.5

Figure 11: Typical soil profile Figure 12: Suction change profile

Step 3: Calculation of ys

Provided the soil suction change profile is known, the characteristic ground movements ys can
be calculated by Equation 1. The overall characteristic surface movement is the sum of the
movement of each of soil layers as given in Table 5

Table 5: The ys calculation for Braybrook


Soil Depth Cracked Iss Z Ipt Δu Average Δu Δz ys
α
Layer (m) Zone (%) (m) (%) (log10(kPa)) (log10(kPa)) (m) (mm)
0 - 1 1.20
1 Yes 4.53 4.53 0.95 0.9 38.68
0.9 - 1 0.70
0.9 - 1 0.70
2 No 5.52 7.18 0.39 1.1 30.84
2.0 2 1.6 0.08
2.0 2 1.6 0.08
3 No 4.90 7.77 0.04 0.15 0.49
2.15 2.15 1.57 0.00

Σ 70
Vol. 22 [2017], Bund. 10 3960

Step 4: Site classification according to AS2870


The calculated characteristic ground movements ys is 70 mm for this site. The site classification
for reactivity following Australian Standard AS2870 is Class H2 (i.e. very highly reactive, Table 1).

CONCLUSION
Expansive soil is widely distributed all over Australia. The ground movement caused by
expansive soil is of great concern not only to the property owners but also to design engineers and
builders. Since 1986, the residential footing design and construction in Australia has been guided by
a national standard, AS2870. In this paper, the current footing design and construction practice in
Australia are reviewed. A worked example for the site classification is also provided to help readers
gain a better understanding of the Australian approach of residential footing design.

AS2870 adopted Thornthwaite Moisture Index (TMI) to account for the impacts of climate
change on footing design. The practice of using TMI to infer the depth of design suction change Hs
required for routine footing design has been widely adopted by geotechnical engineers with success
over the past two decades. To assess climatic variations for Melbourne and Geelong (two largest
cities by population in the State of Victoria), TMI indices have been calculated and analyzed for two
consecutive 30-year periods. The results show that there are demonstrable linear relationships
between annual TMI and rainfall for both cities although there are some variations about the trend
line. The increase in TMI values suggests that both cities have experienced the drying climate. This
implies that residential slabs are likely to experience greater ground movements due to the greater
depth of Hs, which in turn may result in a higher incidence of slab edge heave and an increase in the
occurrence of distortion of residential buildings built on expansive soils.
Vol. 22 [2017], Bund. 10 3961

REFERENCES
1. Aitchison, G.D. and Richards, B.G. (1965) A broad-scale study of moisture conditions
in pavement subgrades throughout Australia: Parts 2 & 3. Moisture Equilibria &
Moisture Changes in Soils Beneath Covered Areas. Butterworths, Sydney.

2. Aitchison, G.D. and Woodburn, J.A. (1969) Soil suction in foundation design.
Proceedings of 7th International Conference on Soil Mechanics and Foundation
Engineering, Mexico City.

3. AS 1289.7.1.1 (2003) Soil reactivity tests -Determination of the shrinkage index of a


soil - shrink-swell index, Standards Australia.

4. AS 1289.7.1.2 (1998) Soil reactivity tests - Determination of the shrinkage index of a


soil - loaded shrinkage index, Standards Australia.

5. AS 1289.7.1.3 (1998) Soil reactivity tests - Determination of the shrinkage index of a


soil - core shrinkage index, Standards Australia.

6. AS 2870 (1996) Residential Slab and Footings, Standards Australia.

7. AS 2870 (2011) Residential Slab and Footings, Standards Australia.

8. Chan, I. and Mostyn, G. (2008) Climatic factors for AS2870 for the Metropolitan
Sydney area. Journal of Australian Geomechanics. Vol. 43(1), pp 17-28.

9. Chan, I. and Mostyn, G. (2009) Climatic factors for AS2870 for New South Wales.
Journal of Australian Geomechanics. Vol. 44(2), pp 41- 46.

10. Chow, V.T. (1964) Handbook of Applied Hydrology. McGraw-Hill, New York, USA.

11. Coffey and Partners Pty. Ltd. (1985) Sydney swelling soils study: analysis of data.
Report to Builders Licensing Board, Report No. S7032/2-AD.

12. Delaney, M. G., Li, J. and Fityus, S.G. (2005) Field Monitoring of Expansive Soil
Behaviour in the Newcastle-Hunter Region. Journal of Australian Geomechanics, Vol.
40 (2), pp. 3-14.

13. Er, Y. and Rifat, B. (2014) Evaluation of climatic factors for the classification of
oklahoma pavement regions. Geo-Congress 2014 Technical Papers. pp 4037-4046.

14. Fityus, S. G. (1996) The effect of initial moisture content and remoulding on the
shrinkswell index, Iss. Proceedings, 7th ANZ Conference on Geomechanics, IE
Australia, Adelaide. pp 388-393.

15. Fityus, S. G., Walsh, P.F. and Kleeman, P. W. (1998) The influence of climate as
expressed by the Thornthwaite index on the design depth of moisture change of clay
Vol. 22 [2017], Bund. 10 3962

soils in the Hunter Valley. Conference on Geotechnical Engineering and Engineering


Geology in the Hunter Valley, Conference Publications, Springwood, Australia. pp 51-
265.

16. Fityus, S. G. and Buzzi, O. (2008) On the use of the Thornthwaite Moisture Index to
infer depths of seasonal moisture change. Journal of Australian Geomechanics, Vol. 43
(4), pp 69-76.

17. Fox, E. (2000) A climate-based design depth of moisture change map of Queensland
and the use of such maps to classify sites under AS 2870-1996. Journal of Australian
Geomechanics. Vol. 35, No. 4, pp 53-60.

18. Fox, E. (2002) Development of a map of Thornthwaite Moisture Index Isopleths for
Queensland. Journal of Australian Geomechanics. Vol. 37, No. 3, pp 51-59.

19. Grismer, M.E., Orang, M., Snyder, R. and Matyac, R. (2002) Pan evaporation to
reference evapotranspiration conversion methods. J. Irrig. and Drain. Engrg. Vol.
128(3), pp 180 -184.

20. Holland, J.E. (1981) The Design, performance and repair of housing foundations,
Swinburne Institute of Technology, Victoria, Australia.

21. Jewell, S.A. and Mitchell, P.W. (2009) The Thornthwaite Moisture Index and seasonal
soil movement in Adelaide, Australian Geomechanics. Vol. 44 (1), pp 59-67.

22. Karunarathne, A.M.A.N., Gad, E.F., Disfani, M.M., Sivanerupan, S. and Wilson, J.L.
(2016) Review of calculation procedures of Thornthwaite Moisture Index and its
impact on footing design. Journal of Australian Geomechanics, Vol. 51(1), pp 85-95.

23. Li, J. and Cameron, D. A. (2002) A case study of a courtyard house damaged by
expansive soils. Journal of Performance of Constructed Facilities, ASCE, Vol. 16 (6),
pp 169-175.

24. Li, J. and Sun, X. (2015) Evaluation of changes of Thornthwaite Moisture Index in
Victoria. Journal of Australian Geomechanics, Vol.50 (3), pp 39-49.

25. Li, J. and Zhou, A. N. (2013) The Australian approach to residential footing design on
expansive soils. Applied Mechanics and Materials, Vol. 438439, pp 593-598.

26. Li, J., Cameron, D. A. and Ren, G. (2014) Case study and back analysis of a residential
building damaged by expansive soils. Computers and Geotechnics, Vol. 56, pp 89-99.

27. Li, J., Zou, J., Bayetto, P. and Barker, N. (2016) Shrink-swell index database for
Melbourne, Journal of Australian Geomechanics, Vol. 51 (3), pp. 61-76.
Vol. 22 [2017], Bund. 10 3963

28. Lopes, D. and Osman, N.Y. (2010) Changes of Thornthwaite’s total moisture indices in
Victoria from 1948-2007 and the effect on seasonal foundation movement. Journal of
Australian Geomechanics, Vol. 45 (1), pp 37-48.

29. McKeen, R.G. and Johnson, L.D. (1990) Climate-controlled soil design parameters for
mat foundations. Journal of Geotechnical Engineering, Vol. 116 (7), pp 1073 -1094.

30. Mitchell, P.W. (2012) Footing design for tree effects considering climate change.
Proceedings 2012 ANZ Geomechanics Conference, Melbourne. No. 1.3.11, pp 290-
295.

31. Mitchell, P.W. (2013) Climate change effects on expansive soil movements.
Proceedings of the 18th ICSMGE, Paris.

32. Perera, Y. Y., Zapata, C. E., Houston, W. N. and Houston S. L. (2004) Long term
moisture conditions under highway pavements. Geotechnical Engineering for
Transportation Projects. Proceedings of Geo-Trans 2004. ASCE Special Publication
126.

33. Richards, B.G., Peter, P. and Emerson, W.W. (1983) The effects of vegetation on the
swelling and shrinking of soils in Australia. Geotechnique, Vol. 33, pp 127-139.

34. Richards, B. G. (1984) Keynote Address. Fifth International Conference on Expansive


Soils, Adelaide.

35. Smith R. (1993) Estimating Soil Movements in New Areas. Seminar-Extending the
code beyond residential slabs and footings, The Institution of Engineers, Australia.

36. Sun, X., Li, J. and Zhou, A.N. (2017) Evaluation and comparison of methods for
calculating Thornthwaite Moisture Index. Journal of Australian Geomechanics, Vol. 52
(2)

37. The AGE, (2011) Owners find homes are cracking under pressure, Viewed 9th July
2015 at http://theage.domain.com.au/home-owning-tips/owners-find-homes-are-
cracking-under-pressure-20111221-1p5or.html.

38. Thornthwaite, C. W. (1931) Climates of North America according to a New


Classification, The Geographical Review, Vol. 21(4), pp 633-655.

39. Thornthwaite, C.W. (1948) An approach toward a rational classification of climate.


The Geographical Review, Vol. 38(1), pp 55-94.

© 2017 ejge
Vol. 22 [2017], Bund. 10 3964

Editor’s note.
This paper may be referred to, in other articles, as:
Xi Sun, Jie Li, and Dr. Gang Ren: “Residential Footing Design on
Expansive Soils: A Review of Australian Practice” Electronic Journal of
Geotechnical Engineering, 2017 (22.10), pp 3939-3964. Available at
ejge.com.

You might also like