Module 3 - Lecture Notes - Additional Reading Material
Module 3 - Lecture Notes - Additional Reading Material
David Oliveira
1 Key references
It is strongly recommended that for this Module 3 you complement your understanding of Rock bolting Design
by reading the Practical Rock Engineering Book by Evert Hoek, freely available at the website below.
https://www.rocscience.com/assets/resources/learning/hoek/Practical-Rock-Engineering-Full-Text.pdf
The following chapters are associated with this module:
https://www.rocscience.com/assets/resources/learning/hoek/Practical-Rock-Engineering-Chapter-5-
Structurally-Controlled-Instability-in-Tunnels.pdf
https://www.rocscience.com/assets/resources/learning/hoek/Practical-Rock-Engineering-Chapter-14-Rockbolts-
and-Cables.pdf
In addition, the papers listed below haven been included in the appendix further discussing the details of some
of the topics presented in this module. They are to be used for educational purpose only and not to be re-
distributed.
Li, C. C. (2017). Principles of rockbolting design. Journal of Rock Mechanics and Geotechnical Engineering
9, pp 396-414.
Oliveira D. (2018). Design of large span tunnels and caverns: back to basics. Invited Keynote: 2018 AGS
Victorian Symposium - Geotechnics and Transport Infrastructure, Melbourne, October, pp.25-34.
Oliveira, D.A.F., Diederichs, M.S. (2017). Tunnel support for stress induced failures in Hawkesbury
Sandstone Tunnelling and Underground Space Technology 64, pp 10-23.
Oliveira, D.A.F., Paramaguru, L. (2016). Laminated rock beam design for tunnel support. Australian
Geomechanics Journal, 51 (4), pp 1-17, December.
Oliveira D. and Pells, P. (2014). Revisiting the applicability of voussoir beam theory for tunnel design in
Sydney. Australian Geomechanics Vol 49(3), pp 29-44, September
Diederichs, M S and Kaiser P K. (1999). Stability of large excavations in laminated hard rock masses: the
voussoir analogue revisited. Int. J. Rock Mech. & Min. Sci., 36, pp 97-118.
Pells, P. J. N. (2002). Developments in the design of tunnels and caverns in the Triassic rocks of the Sydney
region. International Journal of Rock Mechanics and Mining Sciences 39, pp 569-587
1
Ground support for tunnels in rock by Dr. David Oliveira
Appendix (Papers)
2
Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
a r t i c l e i n f o a b s t r a c t
Article history: This article introduces the principles of underground rockbolting design. The items discussed include
Received 17 January 2017 underground loading conditions, natural pressure zone around an underground opening, design
Received in revised form methodologies, selection of rockbolt types, determination of bolt length and spacing, factor of safety, and
3 April 2017
compatibility between support elements. Different types of rockbolting used in engineering practise are
Accepted 4 April 2017
Available online 3 May 2017
also presented. The traditional principle of selecting strong rockbolts is valid only in conditions of low in
situ stresses in the rock mass. Energy-absorbing rockbolts are preferred in the case of high in situ
stresses. A natural pressure arch is formed in the rock at a certain distance behind the tunnel wall.
Keywords:
Rockbolting design
Rockbolts should be long enough to reach the natural pressure arch when the failure zone is small. The
Pressure arch bolt length should be at least 1 m beyond the failure zone. In the case of a vast failure zone, tightly spaced
Bolt length short rockbolts are installed to establish an artificial pressure arch within the failure zone and long cables
Bolt spacing are anchored on the natural pressure arch. In this case, the rockbolts are usually less than 3 m long in
Factor of safety mine drifts, but can be up to 7 m in large-scale rock caverns. Bolt spacing is more important than bolt
length in the case of establishing an artificial pressure arch. In addition to the factor of safety, the
maximum allowable displacement in the tunnel and the ultimate displacement capacity of rockbolts
must be also taken into account in the design. Finally, rockbolts should be compatible with other support
elements in the same support system in terms of displacement and energy absorption capacities.
Ó 2017 Institute of Rock and Soil Mechanics, Chinese Academy of Sciences. Production and hosting by
Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/
licenses/by-nc-nd/4.0/).
Rockbolt is the most widely used support element in support 2.1. Low in situ stress conditions
systems in underground mines and civil tunnels. Rockbolting
design is indeed mainly based on experience and it appears that Rock blocks in the roof of an underground opening are pre-
rockbolting design is simply a business of selection of rockbolt vented to fall as long as a high enough horizontal stress exists in the
types and the determination of bolt length and spacing, but, one rock mass. However, they would fall under gravity in low in situ
essentially uses, either explicitly or implicitly, a methodology in a stress conditions. In locations close to the ground surface, the rock
specific rockbolting design. Attempts are made in this article to mass often contains well-developed rock joint sets. The rock joints
summarise the design principles and methodologies hidden in sometimes are open, which is an indication that the in situ rock
rockbolting practise, which include the relationship between the in stresses are low in the rock mass. The task of rock support in low
situ stress state and rockbolt types, the concept of pressure arch, stress rock masses is to prevent rock blocks from falling. To do so,
design methodologies, determination of bolt length and spacing, the maximum load exerted on the support elements, such as
factor of safety, compatibility between support elements and rockbolts, is the deadweight of the potentially falling block (Fig. 1).
different types of rockbolts. This is a load-controlled condition.
From the point of view of mechanics, the rockbolts must be
strong enough to bear the deadweight of the loosened rock block.
Therefore, use of a factor of safety, defined by the strength of the
E-mail address: Charlie.c.li@ntnu.no. support system and the weight force of the block (i.e. the load), is
Peer review under responsibility of Institute of Rock and Soil Mechanics, appropriate for rock support design in a load-controlled condition.
Chinese Academy of Sciences.
http://dx.doi.org/10.1016/j.jrmge.2017.04.002
1674-7755 Ó 2017 Institute of Rock and Soil Mechanics, Chinese Academy of Sciences. Production and hosting by Elsevier B.V. This is an open access article under the CC BY-
NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 397
3. Design principles
This is essentially the design principle in structure mechanics,
which states that the load applied to a structure should not be
3.1. Natural pressure arch
higher than the strength of the structure, i.e. the strength-to-load
ratio that is called the factor of safety, should be larger than 1.
Geological exploration drilling was once carried out in a mine
This principle is valid for ground constructions where the total load
drift, excavated 5 years previously, at a depth of 1000 m. The mine
on the construction structures is usually known or easily found out.
drift was parallel with the strike of the tabular ore body and the
In shallow underground openings, this principle is also valid since
boreholes were drilled in the wall of the drift on the side of the ore
the maximum load on the rock support system is the deadweight
body that was approximately 150 m apart from the drift. The frac-
force of loosened rock blocks.
ture logging on the cores provided information on the distribution of
the secondary stresses in the rock surrounding the drift. Fig. 3 shows
2.2. High in situ stress conditions the fracture patterns in the cores taken from a horizontal borehole.
The fracture intensity in the cores varies along the borehole. The
The author observed in a deep metal mine that the number of cores are small pieces with a low value of rock quality designation
geological discontinuities in the rock mass became less and the (RQD) in the zone from the wall to a depth of 2.1 m (Zone I). The
discontinuities were less opened in depth. For instance, at a depth fracture surfaces in this zone are yellow coloured, indicating that
of 1000 m, it was observed that all of the few discontinuities they were probably created when the drift was excavated a few
exposed on an excavation face were completely closed. Therefore, it years earlier. The cores are disked in the zone from 2.1 m to 8.5 m
can be said that the rock mass quality is improved at depth because (Zone II). The fractures in this zone are fresh and perpendicular to
of the reduction in the number of geological discontinuities. the core axis. It can be said with confidence that they were created
However, the in situ rock stresses increase with depth. At depth, the during core drilling. Zone II can be further divided into two sub-
major instability issue is no longer fall of loosened rock blocks but zones. In Zone IIa, the core disking is so severe that the disks are
rock failure caused by stress. High stresses could lead to two con- tightly spaced. The disk thicknesses are obviously larger in Zone IIb
sequences in underground openings: large deformation in soft and than in Zone IIa. Zone III is from 8.5 m to the end of the borehole at
weak rock and rockburst in hard and strong rock (Fig. 2). It was the depth of approximately 180 m. The discontinuities in this zone
observed in some metal mines in Sweden that strain burst usually are believed to be mainly of geological origin. The RQD of the cores
occurred below a depth of 600 m and became intensive below in Zone III is significantly higher than the other two zones, which
1000 m. Rock failure is unavoidable in high stress conditions. The implies that Zone III is out of the disturbance distance of the drift. On
task of rock support at depth is not to equilibrate the deadweight the basis of the variation of the fracture intensity, it is inferred that
force of loosened blocks but to prevent the failed rock from disin- Zone I was the failure zone, where the rock failed either in shear or
tegration. In high stress rock masses, the support system must be in tension and the tangential stress was partially reduced, while the
not only strong but also deformable in order to deal with either tangential stress in Zone II was elevated but the rock had not yet
stress-induced rock squeezing in soft and weak rock or rockburst in fractured after excavation of the drift. Zone II was the position of the
hard and strong rock. natural pressure arch that carried the ground pressure and func-
tioned as a protection shield over the drift.
2.3. Suitable rockbolt types To illustrate the failure zone surrounding an underground
opening, numerical modelling was conducted for a horseshoe-
The suitable types of rockbolts for a given rock mass are asso- shaped tunnel of 6 m in width and 6 m in height, excavated in a
ciated with the loading condition in the rock mass. In the case of a rock mass subjected to hydrostatic in situ stresses. The in situ
398 C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
Fig. 2. Loading conditions to rockbolts in high stress rock masses. (a) Rock squeezing, (b) strain burst, and (c) fault-slip burst.
Fig. 3. Cores drilled in the wall of a mine drift at a depth of 1000 m (Li, 2006a).
stresses are assumed to be s1 ¼ s2 ¼ s3 ¼ 30 MPa in the simulation underground opening, where the tangential stresses are signifi-
and the rock mass obeys the Mohr-Coulomb failure criterion with cantly elevated. This is the so-called natural pressure arch, sketched
cohesion c ¼ 5 MPa and internal friction angle f ¼ 35 . The in Fig. 5. The concept of the natural pressure arch was used for rock
constitutive model of the material is elastic and perfectly plastic, i.e. support design by, among others, Wright (1973), Krauland (1983)
the residual strength of the material is equal to the peak strength. and Li (2006b).
Fig. 4 shows the distribution of the major principal stresses that are
oriented approximately parallel with the tunnel walls and roof, i.e. 3.2. Design methodology
in the tangential directions, after excavation. The immediate sur-
rounding rock, approximately 2 m deep in the walls, fails. Beyond Rock support refers in general to any measure aiming to stabilise
that depth, the rock is still intact but the tangential stress has been rock masses by using support elements. Support elements may be
elevated somewhat, depending on the distance to the tunnel wall. It rockbolts, cables, meshes, straps, lacing, shotcrete (i.e. sprayed
reaches its maximum at a depth of about 3 m and then gradually concrete), thin liners, steel sets, shotcrete arches and cast concrete
drops to the in situ stress level (30 MPa) in locations far away from lining. A support system provides three primary functions: rein-
the tunnel. The rock portion within which the tangential stress is forcement, holding and retention (Kaiser et al., 1996). Reinforce-
significantly elevated carries the majority of the ground pressure ment refers to strengthening of the rock mass; holding to the
and forms a protection shield, i.e. a pressure arch, around the suspension of potentially loosened blocks; and retention to
tunnel. confinement of the exposed rock surfaces. Each support element
Based on the core logging shown in Fig. 3 and the numerical may perform one or more of the three primary functions. Rein-
modelling shown in Fig. 4, it can be deduced that a pressure arch (or forcement is usually achieved by installing rockbolts systematically.
ring) exists at a certain depth of the rock surrounding an The increase in the rock strength due to bolting is very limited.
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 399
Fig. 4. Distribution of the major principal stresses in the rock surrounding a tunnel. The crosses and circles mark the zone of rock failure.
Fig. 5. A sketch illustrating the natural pressure arch surrounding an underground opening.
Assume that the load capacity of a rockbolt is 200 kN and rockbolts mainly achieved by shotcrete, mesh or other types of thin liners laid
are installed with a pattern of 1 m 1 m. The maximum confining on the rock surface in mines. In civil tunnels, the allowable rock
pressure that the rockbolts can provide is 0.2 MPa. The increase in deformation is much smaller than that in mines. Therefore, heavy
the rock strength by this confining pressure, according to the external support structures such as steel sets, concrete arches and
MohreCoulomb criterion, may be in the range of 1e2 MPa, which is even cast concrete lining are applied to restrain wall deformation.
significantly lower than the inherent strength of the rock mass. The Those structures are set up on rock surfaces, but they are similar to
essential function of bolting is to keep the fractured rock together cables installed within the rock mass in the sense that they provide
to form a pressure arch around the opened space. In other words, a holding function. A rock support system may be composed of one,
the bolts help the rock to strengthen and support itself. Rockbolts or more than one, of the following support layers, depending on the
also provide a holding function to loosened blocks and fractured loading condition and the extent of rock failure (Li, 2012):
rock. In the case of a large failure zone, rockbolts may be entirely
located within the failure zone. The use of long cablebolts is an (1) Layer 1 e Bolting: Rockbolts are installed sporadically or
option to provide an effective holding function. Surface retaining is systematically.
400 C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
(2) Layer 2 e Surface retaining: Retaining elements like meshes, spaced ductile rockbolts “Split sets” or point-anchored resin bolts
straps, lacing, thin liners, shotcrete and cast concrete lining of 2.4e3 m in length. The bolt-reinforced zone is then nailed to the
are installed on the rock surface. competent strata behind the failure zone with cables (see Fig. 7).
(3) Layer 3 e Cable bolting: Single- or multi-strand cables are The tightly spaced rockbolts help the fractured rock to build up an
installed into the competent rock behind the failure zone. artificial load-bearing arch and the cablebolts integrate the arch
(4) Layer 4 e External support: Structure elements, including with the deeply located stable strata. The rock surface is retained
steel sets, concrete arches, invert, cast concrete lining and with mesh, straps and mesh shotcrete. The Agnew gold mine is
thick shotcrete liners, are set up in tunnels. located in Western Australia. Rock squeezing is the major instability
problem in the mine below the depth of 500 m because of the weak
In shallow tunnels, or in the case of a small failure zone, the task footwall rocks, ultramafic conglomerates, chlorite and talc. The
of the rock support is to hold the loosened or failed rock blocks. wallewall convergence of a 5 m 5 m mine drive reached 400 mm
Support layer 1, i.e. spot bolting or sparsely spaced pattern bolting, in 5 months. The rockbolts used in the mine were either Split sets or
may be good enough to stabilise the rock mass (Fig. 6a). The bolts point-anchored resin Posimix bolts, 2.4e3 m in length and 1 m in
should be installed into the natural pressure arch behind the failure spacing. The twin-strand cablebolts used were 6 m in length and
zone. In poor rocks, the failure zone may be so extensive that 2 m in spacing.
rockbolts cannot reach the natural pressure arch. A support system In burst-prone rock, the support system is composed of 2.4-m or
composed of layers 1 and 2 is then needed (Fig. 6b). Layer 1, in this 3-m long energy-absorbing rockbolts and meshes or fibre/mesh-
case, must be tightly spaced pattern bolting, which builds up an reinforced shotcrete. It is claimed that such a system can sustain
artificial pressure arch in the failure zone. The artificial pressure arch an ejection velocity of 5 m/s and absorb a dynamic energy of 35 kJ/m2
forms a protection shield over the opening. In extremely poor rock, (Slade and Ascott, 2007).
support layer 3, cable bolting, needs to be added to the system. The
1-2-3-layer support system is often used in deep mines and un- 3.2.2. Canadian and Scandinavian methodology
derground caverns of large span, such as underground machine The methodology of rockbolting in Canada and Scandinavian
halls in hydropower plants. This support system is characterised by countries is to integrate the failed rock with short bolts in
its flexibility in order to adapt to the prevailing rock condition. In conjunction with meshes or/and fibre/mesh-reinforced shotcrete
civil tunnels, cable bolting is less used than in mining excavations. (Fig. 8).
Instead, external support structures (layer 4), such as steel sets and In Canadian metal mines, the types of rockbolts are 2.4-m or 2.1-
concrete lining, are employed to provide the holding function. A 1- m long fully resin-encapsulated rebar and Split sets. Energy-
2-4-layer support system is preferred in civil tunnels. absorbing rockbolts are added in case of seismic rock conditions.
Pattern bolting plays a crucial role in a support system. Tightly The Craig mine, in the region of Sudbury, Ontario, is characterised
spaced bolts constrain the failed rock so that an artificial pressure by the faults going through the ore body, which creates a number of
arch is established in the failure zone. The load-bearing capacity of fault-slip seismic events in the mining operation areas below the
an artificial pressure arch was visually demonstrated by Lang depth of 1600 m. In 5-m span drives, the 1.8e2.4 m long fully resin-
(1961) in the 1960s and also recently by Hoek (2007). Li (2006b) encapsulated rebar bolts were used together with surface retaining
reported an example of applying the concept of an artificial pres- support elements (meshes, straps, shotcrete, etc.). In burst-prone
sure arch for rock support design. areas, rebar bolts and modified cone bolts were installed plus 6-
m long cables.
3.2.1. Australian methodology In Swedish mines, fully encapsulated rebar bolts with cemen-
In squeezing rock, the methodology of rockbolting in Australian titious grout are most often used. Split sets are seldom used in
mines is to reinforce the failure zone of the rock using tightly Scandinavian mines, but energy-absorbing rockbolts, such as D-
Fig. 6. Principles of rockbolting in different rock conditions of rock failure: (a) for a limited failure zone, and (b) for a vast failure zone.
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 401
Fig. 8. The Canadian and Scandinavian methodology of rockbolting in deep metal mines.
bolts, have been used for dynamic rock support, for instance, in the absorbing rockbolts and surface retaining elements. It is thought
iron ore mines in Sweden. The surface retaining liners are usually that the kinetic energy in a rockburst event is partially absorbed by
steel-fibre-reinforced shotcrete, but in burst-prone conditions, wire the rockbolts and partially dissipated by fragmentation of the rock
meshes are laid on the top of the steel-fibre-reinforced shotcrete. contained by the surface retaining elements. Mesh and lacing are
The bolt length is typically 2.7e3 m and the bolt spacing is 1e1.5 m. often used in South African dynamic support systems.
The Kristineberg metal mine in Sweden operates the mining ac- In South Africa, mine drifts excavated in high stress rock are
tivity in depths below 1000 m at present. The mine is subjected to typically supported with energy-absorbing rockbolts (e.g. cone
rock squeezing in zinc-and-lead ore bodies owing to the chlorite/ bolts and Durabar) and ductile bolts (e.g. split set and cables). The
talk-rich rocks, but was subjected to strain burst in the hard primary bolting method is a ring of 1.2-m long bolts and the sec-
quartzite of the gold ore bodies. The rockbolts used in the mine ondary bolting is a ring of 2.4-m long bolts plus lacing, meshing
have been 2.7-m long fully cement-grouted rebars with a bolting and 50 mm thick steel or polyester fibre shotcrete (Fig. 9). The
pattern of 1.2 m 1.2 m. bolting pattern is typically 1 m 1 m. The Mponeng gold mine in
South Africa operates its mining activity at depths below 2500 m.
3.2.3. South African methodology The dominant rock is quartzite that is burst-prone under high
In burst-prone deep mines in South Africa, the methodology of ground pressure. Rockbolt is only one of a number of ground
rockbolting is to dissipate the released kinetic energy with energy- support elements used in the mine. Split sets, 1.2 m long, are used
402 C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
in mining stopes and Durabars, 2.4e3.6 m long, are used in In practise, the bolt pattern in systematic bolting is such that the
transport drives. in-row spacing and the distance between rows are equal. The bolt
spacing, s, is recommended to be in the range from 1 m to 2.5 m.
3.3. Bolt length and spacing However, rock joint spacing should be also taken into account
when the bolt spacing is determined. A rule of thumb is to set the
Determination of bolt length and spacing has been a topic for bolt spacing equal to 3e4 times the mean joint spacing in the case
discussion probably since rockbolts were first used for ground of a mean joint spacing in the range of 0.3e1 m, i.e.
support in underground excavations (e.g. Panek, 1964; Coates and
Cochrane, 1970; Lang, 1972; Barton et al., 1974; Schach et al., s ¼ ð3 4Þe (4)
1979; Farmer and Shelton, 1980; Crawford et al., 1985; Stillborg,
1994). Choquet and Hadjigeorgiou (1993) provided a review on where e represents the mean joint spacing.
this topic in their paper on the design of ground support. The In the case of a vast failure zone (Fig. 6b), the Norwegian Road
following presented are the principles for the determination of bolt Authority recommends the use of relatively short tensioned rock-
length and spacing that are used in the practise of rockbolting to bolts to establish an artificial pressure arch in the failure zone (see
date. From the point of view of operation, the bolt length should be Fig. 10b). The bolt length is still estimated using Eq. (3), but the bolt
less than half of the opening height for roof bolts and half of the spacing is recommended to be smaller than 3 times the mean joint
span for wall bolts in order to avoid installation difficulties: spacing, i.e.
Lb df þ 1 (2) The bolt length is usually short, 2e3 m, in this type of rock-
bolting. The thickness of the interaction zone is required to be at
where df is the depth of the failure zone. In the case of a vast failure least 0.5Lb in order that a strong enough artificial arch can be
zone (Fig. 6b), the bolt length is short, varying from 2 m to 3 m, but established in the broken rock. This requirement leads to a bolt
its upper limit is governed by Eq. (1). spacing that should be less than half of the bolt length, i.e.
For tunnels excavated in moderately jointed hard rock masses,
the Norwegian Road Authority proposed the following formula to s Lb =2 (7)
determine the length of un-tensioned bolts in the central section of In the design stage of an underground rock excavation, bolt
the tunnel for the purpose of suspending the failure zone on the length and spacing are often determined with the help of empirical
natural arch (Statens vegvesen, 2000) (see Fig. 10a): methods recommended in various rock mass classification systems.
In the Q rock mass classification system (Barton et al., 1974), the bolt
Lb ¼ 1:4 þ 0:184B (3)
length and spacing can be found in a chart based on the Q-value of
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 403
uc
FS ¼ (9)
ueq
It must be pointed out that the critical displacement uc is
difficult to be quantified even with the help of numerical model-
ling. Beyond displacement uc, the rock mass becomes unstable and
calculation iterations would become non-convergent in numerical
modelling. In engineering practise, there usually exists a
maximum allowable displacement from the point of view of
operation. For example, the radial displacement of a TBM (tunnel
Fig. 10. Rockbolting methods in two different rock conditions: (a) suspension of the
failure zone to the natural arch and (b) establishment of an artificial arch within the
boring machine) tunnel usually is not allowed to be larger than
failure zone. Modified after Stillborg (1994). 150 mm in order to avoid jamming of the TBM head. Thus, there is
another factor of safety for the operation, denoted as FSop, which is
calculated as
the rock mass and a geometrical parameter called the equivalent umax
FSop ¼ (10)
dimension (Barton and Grimstad, 2014). The equivalent dimension ueq
is defined by the span of the excavation and a coefficient describing
the intended use of the excavation (road tunnel, underground where umax is the maximum allowable displacement. In addition to
station, etc.). In the rock mass rating (RMR) system by Bieniawski the factors of safety for stability and operation, it is also required
(1989), bolt length and spacing, as well as other types of support that the rock displacement at equilibrium, ueq, must be smaller
measures, are empirically recommended in a table for five classes than the ultimate displacement uult of the rockbolt to avoid pre-
of rock mass quality. mature failure of the rockbolt. To the end, the items in the criteria
for a rock support system using rockbolts are as follows:
Fig. 12. The ground response curve (GRC) and the support characteristic curve of yield rockbolts.
Fig. 14. Sketches illustrating incompatible rock support systems: (a) in civil tunnelling and (b) in mining.
be equal. Given r ¼ 2700 kg/m3, g ¼ 10 m/s2 and f ¼ 35 , the critical reinforcement force contributed by the bolts is T ¼ St where t is the
dip angle is related to the tangential rock stress sq and the block force in a single rockbolt. All forces exerted on the block are illus-
height h, as shown in Fig. 18. The critical dip angle decreases with an trated in Fig. 19b. They are the gravitational force acting on the
increase in the tangential rock stress and approaches 55 , which is block, Wg, the total bolt force, T, the normal reaction force on the
the critical dip angle without consideration of gravity. The critical dip sliding plane, N, and the shear resistant force on the sliding plane, R.
angle increases with a decrease in the tangential rock stress, All forces are in equilibrium in all directions in the critical state at
implying that only steeply dipped wedge blocks could be stabilised which shear failure occurs along the sliding plane. By equilibrating
by the friction on the block sides in low stress conditions. The critical the forces, it is obtained that the normal force is expressed as
dip angle changes abruptly with a small change in the tangential rock
stress below 1 MPa, but it is not very sensitive to the rock stress N ¼ Wg cos j þ T sin a (17)
above 2 MPa. As shown by the three curves corresponding to
different block heights, the critical dip angle also increases with the where j is the dip angle of the sliding plane. The shear resistant
block size, particularly when the tangential rock stress is lower than force is expressed, with an assumption that the MohreCoulomb
1 MPa. The critical dip angle increases slightly with the block size criterion prevails along the sliding plane, as
when the rock stress is higher than 2 MPa. This implies that the
critical dip angle is not sensitive to block size when the tangential R ¼ cA þ ðWg cos j þ T sin aÞtan fa (18)
rock stress is high.
where A is the base area of the sliding plane. The driving shear
4.1.2. Bolting force, D, is obtained as
The load on rockbolts that are used to stabilise a rock block in
the roof is the deadweight force of the block (see Fig. 1). The D ¼ Wg sin j T cos a (19)
number of bolts needed, Nbolt, is approximately calculated as The factor of safety of the wedge block is defined as the ratio of
the shear resistant force to the driving shear force, i.e.
Wg
Nbolt ¼ FS (16)
Pult
R cA þ ðWg cos j þ T sin aÞtan fa
FS ¼ ¼ (20)
where W is the deadweight force of the block, and Pult is the ulti- D Wg sin j T cos a
mate load of the rockbolt. In the case of fully grouted rockbolts, the A factor of safety less than 1, i.e. FS 1, means that sliding oc-
embedment length of the bolts in the stable formation must be at curs, while the block is stable when FS > 1. A factor of safety in the
least 1 m. range of 1.5e2 is usually used for rockbolting design.
The bolt force T contributes to an increase in the normal force
4.2. Wedge block in wall and a component of the shear force. The increase in the normal
force is always positive in enhancing the frictional resistance of the
The principle of stabilising a wedge block in the wall was pre- sliding plane, but the contribution of the bolt force to the shear
sented, for instance, by Hoek and Brown (1980) and Harrison and force is either positive or negative, depending on the installation
Hudson (2000). It is demonstrated in this section through the angle, a. There exists a theoretical critical installation angle,
example illustrated in Fig. 19a. Assume that a wedge block is denoted as ac, at which the bolts most effectively reinforce the
formed in the wall and it tends to slide along the lower disconti- block. Let FS ¼ 1 in Eq. (20), representing the equilibrium state
nuity under gravity. The block is stabilised with bolts which are when shear failure is initiated along the sliding plane. The bolt force
installed with an angle of a to the discontinuity plane. The total is expressed at this moment as
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 407
ac ¼ fa (22)
In other words, the reinforcement effect of the rockbolts is op-
timum when they are installed at an angle of a ¼ fa.
where smax is the maximum ground pressure that the pressure arch
can bear, and sc is the uniaxial compressive strength (UCS) of the
bolt-reinforced rock party. The coefficient k is proportional to the
moment arm length in the pressure arch. Wright (1973) found that
k is approximately 0.9 based on back-calculations of his experi-
Fig. 16. A kinematic wedge bounded by three planar geological discontinuities and a mental data.
horizontal overhanging rock face.
90°
Fig. 18. The critical dip angle versus the tangential rock stress in the rock for three different block sizes.
Fig. 19. Use of rockbolts to stabilise a wedge block that tends to slide along the lower discontinuity plane. (a) The block and the rockbolts, and (b) the forces on the wedge block.
of 0.6 MPa in the residual strength could significantly improve the post- cavern. The pillar was subjected to extension fracture during
failure behaviour of the pillar. excavation. Cablebolts with a load capacity of 1 MN were installed
Fig. 23 shows the tieback bolting of a 6 m 8 m (width depth) across the pillar with a spacing of 2 m. They were pre-tensioned to
pillar between two niches excavated in the wall of a hydropower 400 kN after installation. A concrete lining of 300 mm in thickness
was cast on both sides of the pillar to improve the load transfer
from the cablebolts to the pillar. Strong square plates of
Pressure arch 200 mm 200 mm were attached to the cablebolts. The maximum
confining pressure that the cablebolts can provide is approximately
1 MN/(2 m 2 m) ¼ 0.25 MPa. The increased strength of the pillar
is estimated to be 0.75 MPa, corresponding to an increase of the
load capacity of 36 MN for the 6 m 8 m pillar (assuming an in-
ternal friction angle of 30 ).
the bolt spacing. The required ultimate load capacity of the 4.6. Rockbolting in large-scale caverns
rockbolt is
Rock failure may extend to a significant depth after excavation
of a large-scale underground cavern. Rockbolts and cablebolts play
Pult ¼ FSðlsC rgÞ (25)
crucial roles in the support system in this case because other
where l is the thickness of the weak layer, and C is the bolt spacing support elements, such as lining, shotcrete and mesh, can only
between rows. The minimum bolt length, Lmin, is passively respond to the rock deformation and provide very
limited effective support to the rock mass. When the rock mass
quality is poor and the in situ rock stresses are high, the size of the
Lmin ¼ l þ anchorage length (26) failure zone could be so vast that it is beyond the bolt-reinforced
The anchorage length in the stable stratum should be at least zone and the reinforced rock party continues to move toward
1 m in the case of fully encapsulated bolting. The principle is that it the opening. The principle of rockbolting in this situation is to
must be longer than the critical embedment length with a factor of reinforce the rock with tightly spaced and fully grouted rockbolts
safety of 2e4. to a relatively shallow depth (3e7 m) in combination with long
cablebolts (10e25 m). The short rockbolts aim to build up a bolt-
rock ‘shield’ surrounding the opening and the long cablebolts
suspend the ‘shield’ to competent and stable rock formations
Fig. 23. Tieback bolting in a hydropower cavern. Some of the pillar-through cablebolts
Fig. 22. Tieback bolting. are shown in the picture (Photo by J. Mierzejewski).
410 C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
behind the failure zone. It is required that the cablebolts must be approximately dip toward the upstream with an angle of 30 . The
able to tolerate a significant displacement in order that they in situ horizontal stress in the rock mass is slightly higher than
function properly without premature failure. A common practise the vertical stress. The final wallewall convergence of the cavern
to enhance the deformability of a cablebolt is to de-bond its middle is expected to be 200e300 mm. The 7-m long short rockbolts,
portion from the grout with PVC pipes or other types of soft with load capacity of 300 kN, are fully grouted in the boreholes.
materials. The short bolts are spaced 1 m 1 m in the crown and
Fig. 25 shows the rockbolting design for a hydropower cavern 1.3 m 1.3 m in the walls. The load capacity of every cablebolt is
located at a depth of approximately 400 m. The cavern is 25 m in 1 MN. In every bolting profile, the cablebolts are spaced
span, 45 m in height and 120 m long. The main rock types are 4 m 4 m in the crown with the three in the middle being 15 m
sandstone (UCS of 80e150 MPa), siltstone (UCS of 40e60 MPa) long and the rest 10 m long. The cablebolts in the walls are 15 m
and mudstone (UCS of 20e50 MPa). The bedding planes long and spaced 5 m 5 m. Three 20-m cablebolts are installed in
Fig. 25. An example of rockbolting and cablebolting in a large-scale underground hydropower cavern.
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 411
the lower portion of the wall on the downstream side because of Table 1
the risk that the rock mass there may be weakened by excavations Specification of forepoles (Bang, 1984; Ocak, 2008; Volkmann and Schubert, 2009;
DSI, 2015; Hoek, 2015).
nearby.
Type of forepoles Diameter (mm) Length (m) Spacing (m) Angle (o)
A B
qs qs
la la
Lr Lr
(a) (b)
Fig. 29. Forepole umbrellas. (a) Without supporting arch in between and (b) with a supporting arch in the middle.
bending nature of the poles (Fig. 28). The uniformly distributed deflectable in the section close to the near end of the pole. If only
load along a pole is expressed by qs, where q is the ground pressure one arch is set up at the near end of the poles (Fig. 29a), every pole
and s is the spacing between poles. The pole section in area A can be functions as a cantilever beam. The maximum deflection of the
simplified to a cantilever beam loaded by the uniformly distributed poles is expressed, according to Eq. (27), as
load qs. The maximum deflection, dA,m, of this pole section is
expressed, according to beam theory, as
qs
dA;m0 ¼ 5:4 L4r 103 (29)
qs EI
dA;m ¼ 5:4 l4a 103 (27)
EI where Lr represents the round length. If there is an additional
supporting arch in the middle of distance Lr, the span between the
where la is the arch spacing, E is the Young’s modulus of the pole
arches becomes Lr/2. The maximum deflection of the pole section
material, and I is the bending moment of the pole. Every pole
close to the near end of the pole becomes
section between two adjacent arches in area B can be simplified by
an end-fixed beam. The maximum deflection, dB,m, of the fixed-end
beam is expressed as 1
dA;m1 ¼ d (30)
16 A;m0
qs
dB;m ¼ 2:6 l4a 103 (28) In other words, the deflection of the pole with an additional arch
EI
in the middle of the round is reduced to one sixteenth of the
It is seen by comparing the two expressions above that the deflection without the additional arch. This means that the middle
maximum deflection of the pole in area B is approximately one half arch is effective in making the poles stiffer.
of the deflection of the pole section in area A. The pole is thus most The following is an example of spiling adopted for rock support
in squeezing rock in a metal mine. The spiles are either rebars or
self-drilling rockbolts depending on the rock conditions. The
rockbolts are at least 6 m long for an advance length of 4 m and fully
grouted with cement in the boreholes. In the case of using self-
drilling bolts, the bolt must be drilled at least 1 m beyond the
advance face. No matter what type of spiles is used, they must be
longer than the advance length of blasting.
Installation of the spiles is depicted in Fig. 30. Spiling holes are
located approximately 1 m above the contour. They are drilled 10 e
15 upward with a spacing of 0.3 m.
Fig. 30. Spiling in a mine drift. Fig. 31. Systematic rockbolting in burst-prone rock conditions.
C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414 413
4.8. Rockbolting in burst-prone rock Bang S. Limit analysis of spiling reinforcement system in soft ground tunneling.
Tunnel 1984;3:140e6.
Barton N, Grimstad E. Tunnel and cavern support selection in Norway, based on
Energy-absorbing rockbolts, such as the D-bolt and cone bolt, rock mass classification with the Q-system. Publication No. 23. Norwegian
should be installed in burst-prone areas in order to achieve a Tunnelling Society; 2014. p. 45e77.
satisfactory reinforcement effect. The occurrence of rockburst is Barton N, Lien R, Lunde J. Engineering classification of rock masses for the design of
tunnel support. Rock Mechanics 1974;6(4):189e239.
random both in position and time. Thus rockbolts are usually sys- Bieniawski ZT. Engineering rock mass classifications. New York: Wiley; 1989.
tematically installed, for example, with a spacing pattern s s (see Choquet P, Hadjigeorgiou J. The design of support for underground excavations. In:
Fig. 31). Assuming that the expected ejection depth is t, the bolt Excavation, support and monitoring. Comprehensive rock engineering e prin-
ciples, practice and projects, vol. 4. Pergamon Press; 1993. p. 313e48.
length should be at least 1 m longer than the ejection depth. The Coates DF, Cochrane TS. Development of design specifications for rock bolting from
bolt spacing s is then required according to Eq. (14) to be research in Canadian mines. Research Report R224. Mining Research Centre,
Energy, Mines and Resources Canada; 1970.
1 2Eab Crawford AM, Bray JW. Influence of the in situ stress field and joint stiffness on rock
s2 ¼ (31) wedge stability in underground openings. Canadian Geotechnical Journal
FS t rV 2 1983;20(2):276e87.
Crawford AM, Ng L, Lau KC. The spacing and length of rock bolts for underground
It is crucial that the rockbolts have a strong link with surface openings in jointed rock. In: Einsenstein Z, editor. Proceedings of the 5th in-
retaining elements such as the mesh and straps so that the load on ternational conference on numerical methods in geomechanics. A.A. Balkema;
the surface support elements is transferred to the rockbolts. 1985. p. 1293e300.
DSI. Spiles and forepoling boards. Product catalog. DWIDAG-System International;
2015.
Farmer IW, Shelton PD. Factors that affect underground rockbolt reinforcement
5. Conclusions systems design. Transactions of the Institutions of Mining and Metallurgy
1980;89:68e83.
The strength of rockbolts is the key parameter for rockbolting Harrison P, Hudson J. Engineering rock mechanics. Part 2: illustrative worked ex-
amples. Pergamon; 2000.
design in low stress rock masses. Rockbolts should be deformable in Hoek E, Brown ET. Underground excavations in rock. London: Institution of Mining
addition to the requirement of high strength in high stress rock and Metallurgy; 1980.
masses. In other words, rockbolts should be energy absorbent in Hoek E. Model to demonstrate how bolts work. In: Practical rock engineering; 2007.
Hoek E. Numerical modelling for shallow tunnels in weak rock. 2015. https://www.
squeezing and burst-prone rock conditions.
rocscience.com/documents/pdfs/rocnews/Spring2003/ShallowTunnels.pdf.
There exists a natural pressure arch immediately outside of the Kaiser PK, Tannant DD, McCreath DR. Canadian rock burst support handbook.
failure zone in the rock surrounding an underground excavation. In Sudbury, Canada: Geomechanics Research Center; 1996.
Krauland N. Rockbolting and economy. In: Stephansson O, editor. Rockbolting e
the case of a shallow failure zone, the rockbolts should be long
theory and applications in mining and underground construction. Rotterdam:
enough to reach the pressure arch. In the case of a vast failure zone, A.A. Balkema; 1983. p. 499e507.
short rockbolts are tightly installed to establish an artificial pres- Lang T. Rock reinforcement. Bulletin of the International Association of Engineering
sure arch within the failure zone and long cables are anchored into Geology 1972;9:215e39.
Lang TA. Theory and practice of rockbolting. Transactions of American Institute of
the natural pressure arch. Mining, Metallurgical, and Petroleum Engineers 1961;220:333e48.
Determination of the bolt length and spacing is associated with Li CC. Design principles of rock support for underground excavations. In: Eurock
the methodology of rockbolting. In the case of the anchorage of 2012. Stockholm, Sweden; 2012.
Li CC. Evaluation of the state of stress in the vicinity of a mine drift through core
rockbolts in the natural pressure arch, the bolt length should be at logging. In: Proceedings of the 4th Asian rock mechanics symposium. New
least 1 m beyond the failure zone. In the case of establishing an Jersey, USA: World Scientific; 2006a.
artificial pressure arch, appropriate bolt lengths are approximately Li CC. Rock support design based on the concept of pressure arch. International
Journal of Rock Mechanics and Mining Sciences 2006b;43(7):1083e90.
3 m in mine drifts and up to 7 m in large-scale hydropower caverns. Nomikos PP, Sofianos AI, Tsoutrelis CE. Symmetric wedge in the roof of a tunnel
Bolt spacing is more important than bolt length in this case. The excavated in an inclined stress field. International Journal of Rock Mechanics
principle is that the bolt spacing guarantees that the rockbolts and Mining Sciences 2002;39(1):59e67.
Nomikos PP, Yiouta-Mitra PV, Sofianos AI. Stability of asymmetric roof wedge under
interact with each other. The appropriate bolt spacing is 1 m for 3-
non-symmetric loading. Rock Mechanics and Rock Engineering 2006;39(2):121e9.
m long bolts and less than 1.5 m for 7-m long bolts. Ocak I. Control of surface settlements with umbrella arch method in second stage
The rockbolting design is based on the deadweight force of excavations of Istanbul Metro. Tunnelling and Underground Space Technology
2008;23(6):674e81.
falling blocks and the strength of the rockbolt in low rock stress
Panek LA. Design for bolting stratified roof. Transactions of the Society of Mining
locations. For high rock stresses, one should take into account the Engineers 1964;229:113e9.
portion of the rock-released energy that needs to be taken care by Schach R, Garschol K, Heltzen AM. Rock bolting: a practical handbook. Oxford:
the rockbolts. The maximum allowable displacement and the ulti- Pergamon; 1979.
Schubert W. Recent experience with squeezing rock in Alpine tunnels. In: CUC e
mate displacement capacity of the rockbolt should also be taken rock support in medium to poor rock conditions; 2001.
into account. Shi G, Goodman RE. Key block bolting. In: Rockbolting e- theory and applications in
The rockbolts in a rock support system should be compatible mining and underground construction. Rotterdam: A.A. Balkema; 1983. p.143e67.
Sinha RS. Rock reinforcement. In: Underground structures e design and instru-
with other support elements with respect to displacement and mentation. Amsterdam: Elsevier; 1989. p. 129e58.
energy absorption capacities. Slade J, Ascott B. Impact of rockburst damage upon a narrow vein gold deposit in
the Easter Goldfields, West Australia. In: Challenges in deep and high stress
mining. Australian centre for geomechanics; 2007. p. 247e56.
Conflict of interest Sofianos AI. Stability of wedges in tunnel roofs. International Journal of Rock Me-
chanics and Mining Sciences & Geomechanics Abstracts 1986;23(2):119e30.
Statens vegvesen. Fjellbolting. Håndbok 215. Oslo: Trykkpartner AS; 2000 (in
The author wishes to confirm that there are no known conflicts Norwegian).
of interest associated with this publication and there has been no Stillborg B. Professional users handbook for rockbolting. Clausthal-Zellerfeld: Trans
significant financial support for this work that could have influ- Tech Publications; 1994.
Volkmann GM, Schubert W. Effects of pipe umbrella systems on the stability of the
enced its outcome. working area in weak ground tunneling. In: Sinorock e international sympo-
sium on rock characterisation, modelling and engineering design methods.
International Society for Rock Mechanics; 2009.
References Wright FD. Roof control through beam action and arching. In: SME Mining Engi-
neering Handbook, vol. 1. New York: Society of Mining Engineers; 1973. p. 80e
Aksoy CO, Onargan T. The role of umbrella arch and face bolt as deformation pre- 96. Chapter 13.
venting support system in preventing building damages. Tunnelling and Un- Zhang C, Feng XT, Zhou H, Qiu S, Wu W. Case histories of four extremely intense rock-
derground Space Technology 2010;25(5):553e9. bursts in deep tunnels. Rock Mechanics and Rock Engineering 2012;45(3):275e88.
414 C.C. Li / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 396e414
Dr. Charlie C. Li is professor of rock mechanics for civil and out thorough studies of rock behaviour under compression in laboratory. He made
mining engineering at the Norwegian University of Science numerous field observations of rock failure and responses of rock support elements
and Technology (NTNU) in Norway. Li received his BSc in mines and other types of underground excavations. After a thorough study of the
degree in 1981 and MSc degree in 1984, both in geological performances of rockbolts, he proposed analytical models for the rockbolts currently
engineering, in Central South Institute of Mining and used in rock engineering practise, which have been acknowledged in the circle of
Metallurgy (at present Central South University), and his rock mechanics. Based on the models, Li identified the shortcomings of the conven-
PhD in mining rock mechanics at Lulea University of tional rockbolts and pointed out that rockbolts, as well as other support elements,
Technology (LUT), Sweden, in 1993. After that, he was must be not only strong and but also deformable, i.e. energy-absorbing, in high in
employed as a research associate and then associate pro- situ rock conditions. He invented a new type of energy-absorbing rockbolt, called D-
fessor at LUT until 2000. He worked then in the Kristine- Bolt in 2006. The D-Bolt is as strong as a fully encapsulated rebar bolt but its deforma-
berg mine of Boliden Mineral Ltd., Sweden, as a mining tion capacity is significantly higher than that of the rebar bolt. The D-Bolt is particu-
engineer for 4 years. He has been the professor of rock larly powerful in combating stress-induced rockburst and squeezing. The bolt has
mechanics at NTNU since 2004, in charge of the teaching been used worldwide in many deep mines and also in hydropower projects, for
and research program in the subject of rock mechanics as instance in Sweden, Canada, USA, Chile, Australia and South Africa, to combat insta-
well as the rock mechanics laboratory. He is a member of the Norwegian Academy of bility problems of rockburst. Dr. Li has practical expertise in ground support in difficult
Technological Sciences (NTVA). He is the European Vice-President of the International rock conditions (for instance, rock squeezing and rockburst), stability analysis of un-
Society for Rock Mechanics (ISRM) for the term of office 2015e2019. Prof. Li’s research derground caverns and in situ measurements and interpretation. His current research
interests are in rock failure, stability analysis of underground spaces, ground support interests are on understanding of rockburst and theories and practise of dynamic rock
and application of rock mechanic principles for underground space design. He carried support.
Keynote Address
ABSTRACT
Increased demand to future-proof tunnel projects with respect to traffic has led to the proposal of some very large spans
in recent road tunnel projects in Australia. For example, four lane tunnels are currently under construction in Sydney with
mined spans of approximately 20 m and Y-junction caverns of unprecedented spans for road tunnels in Australia, all with
a requirement for 100-year design life. As these spans are unprecedented in Australian civil tunnels, a direct comparison
with local past experience is not possible and simple extrapolation of precedent designs, although potentially solving the
problem, often result in uneconomical solutions that do not necessarily target the actual failure mechanisms involved in
the excavation of such large spans. International experience could certainly be used but adequate design justification
would still have to be provided. Although there is certainly room for cutting edge innovation, robust solutions can also be
achieved by simply going back to basics. As a result, this paper intends to present and discuss how designs that focus on
first principles and the basic objectives of rock reinforcement may allow for a better understanding of the design
requirements and how to satisfy codes and standards but also provide savings with respect to ground support. The key to
the design involves understanding the failure mechanism that needs to be addressed, its relationship with the different
actions of rock bolting, i.e. suspension/anchorage and/or rock reinforcement and what could be acceptable.
Keywords: large span, rock support design, rock bolt reinforcement, suspension
M4
ZD
y-va
—‘
spans; and (2) provide uneconomical solutions. to approximately one third of the span. On the other
International experience could certainly be used but hand, only a small number of projects with spans greater
fa
adequate design justifications and analysis would still than 20 m exist in Australia and if they are included, the
have to be provided to verify its application locally. global best fit indicates a power curve that deviates from
r“!
the linear trend. Based on such a best fit, a bolt length of
The search for solutions to new problems often target approximately 6 m would be considered a precedent
innovation. However, considering that the new challenge design for tunnels of approximately 20 m span.
described above in fact involves an “old" problem but at
a larger scale, it is considered appropriate to review the For tunnels where instability is generally controlled by
basic design assumptions to find robust solutions in geological structures, the rule of thumb for bolt length
more fundamental design principles. As a result, this approximately equal to one-third of the span is
paper intends to present and discuss how a design that somewhat related to conservative assessments of the
focus on the basic objectives of rock reinforcement may largest possible wedge assuming ubiquitous rock
allow for a better understanding of the design defects and ignoring stress effects. These assessments
7‘!
requirements and still provide savings with respect to generally result in wedge widths that span approximately
ground support. The key to the design involves the entire tunnel as illustrated in Figure 3, and wedge
1
r'1
understanding the failure mechanism that needs to be apex heights approximately equal to one third of the
K
addressed, its relationship with the different actions of span, particularly for tunnels smaller tunnels of, say,
. 1
rock bolting, i.e. suspension/anchorage and/or rock 15 m span or less.
reinforcement and what could be acceptable.
For the flat—roofed tunnels typically excavated in Sydney,
1
r.‘
2 PRECEDENT DESIGN the design approach generally was that proposed by
L
Bertuzzi and Pells (2002) as illustrated in Figure 4. The
L 4
design is no longer governed by structurally controlled
r'fi
The first step in any design is typically a comparison with
precedent experience of what has worked and what has wedges but involves selecting a roof beam thickness
A
not. Therefore, Figure 2 presents a comparison between which would in turn dictate the rock bolt length based on
"fir-“era
L
span and bolt length for several successful projects in deflections that are deemed acceptable. This deflection
Australia prior to 2014. typically varies between 10 mm and 20 mm. It is
LA
important to note that Bertuzzi and Pells (2002)
Significantly experience is observed for tunnels under suggested to add 1 m extra in bolt length to the beam
18m span. For these span tunnels, the bolt length thickness for embedment purposes.
LA
roughly follows a linear relationship with bolt length equal
PM‘
1o /
9 a a a
a /V ,_ ..
x
7 <>—————
1
7’1
y = 0.6308x° 75‘“
.
R2 = 0.8702
Bolt Length (m)
rd
U"
r‘w
4
0
0 5 10 15 20 25 30
Span (m)
——Spani4 (Lang 1958) 2 t 015 x Span (Barton at at 397-5) 14 s 0184 x Span (Barton et al 1995)
o 609a MS 9 Bondr Pump Station at Opera Hourso Car Park
at M2 Motorway A Eastem Distributor 0 Elgas Distrbutor
r Nonnsrdo Storage Tunnel - Poauna A Lane Cove Tunnel
0 Energy Australia Cabte Tunnel A ECRL x Yransgnd Canto Tunnel
x Transgnd Plant Room 0 Buranda Street 0» 80990 Rd
Coal Mnos NSB‘I’ », Brunswrck Street
NST APL Cross CrtyTunneI‘
How Southern Ralway' I Lutwycho Cavern t: Kodron Cavern
o . . o Lino 018w Flt Span'3 Linear arch (E=1 56Pa‘q=5t]<Pa den <20mm)
STEP 1
Define rock beam thickness (i.e. bolt length) using
linear arch theory and an assummed acceptable
Roof wedge
Weight 44.4 t deflection (10-15 mm).
FS = 0
STEP 2
Assess induced stresses within pseudo<equivalent
rock beam.
STEP 3
Estimate excess shear stress within rock beam.
3‘ STEP 5
Floor wedge '
Weight 181 t
Assess rock bolt contribution as reinforcement
Stable element.
For a 20 m span tunnel, the required bolt length would be The precedent design above already provided some initial
approximately 6 m bolt length to achieve a mid-span indication that in large span tunnels, bolt length may not
deflection of approximately 15mm in accordance with follow similar trends to those of the smaller span tunnels.
Bertuzzi and Pells (2002) and assuming a fair quality Two main factors are considered to influence this change:
pseudo-equivalent rock beam with E =1500 MPa. This is stress effects and a change in design focus to
somewhat consistent with the best-fit of all projects shown reinforcement effects rather the suspension of unstable
in Figure 2. wedges.
The limited data beyond 18 m span indicate a distinct 3.1 Stress effects
difference from the smaller tunnels. Most of the larger
span bolt lengths deviate from both the linear-arch theory As previously discussed, for tunnels where instability is
curve and the span/3 linear trend and approximately generally controlled by geological structures, the rule of
approach the empirical rule originally devised during the thumb for bolt length is somewhat related to conservative
Snowy Mountains project in Australia by Lang (1958) for assessments of the largest possible wedge assuming
bolt lengths of span/4 (for spans greater than 18 m) and ubiquitous rock defects and ignoring stress effects. These
those of Barton et al (1974) and Barton et al (1995). The assessments generally result in wedge widths that span
exceptions are the Eastern Distributor cavern with 9 m approximately the entire tunnel and wedge apex heights
long bolts and span up to 24 m and the Lane Cove Tunnel approximately equal to one third of the span. Making
with 7 m bolts for a 22 m span excavation which are in similar assumptions in large tunnels and caverns, very
Sydney. These two projects still follow approximately the large wedges would therefore be found in the roof which
span/3 or linear arch theory curve as proposed by would require excessively long cable bolts.
Bertuzzi and Pells (2002). On the other hand, the other
projects are outside Sydney, namely the Poatina Power Figure 5 illustrates an unstable wedge, i.e. factor of safety
Station cavern in Tasmania with 7 m bolts for an FoS = 0, with an apex height of approximately 12 m within
approximately 29 m span excavation, the Kedron cavern the roof of a 40 m span cavern. Such a wedge would
in Brisbane with 7m bolts for a 26m span and the require cable bolts lengths of approximately 12—13m.
Lutwyche cavern with 6 m bolts for an approximately 26 m Firstly, the designer must question if such a large-scale
excavation. The main difference between the larger wedge is indeed reasonable based on the typical
excavations outside Sydney and the large span tunnels persistence of the controlling discontinuities. Secondly,
excavated in Sydney is that they are generally all fully an important assumption typically ignored may be
arched structures where the tunnels in Sydney are semi- reconsidered which is the positive clamping effect of
flat roof tunnels, i.e. very low arch tunnels. Stress arching horizontal stresses. If a stress state equal to the in-situ
and combined use of passive shotcrete support in arched stress is assumed, despite horizontal stress likely
excavations may allow for shorter bolts than what would increasing due to the excavation, the wedge would show
have been used in semi-flat—roofed tunnels in Sydney a stable behaviour with a FoS = 1.7.
using the Bertuzzi and Pells (2002) approach as depicted
in Figure 2. As a result, such difference must be Hoek (2007) states that this large difference in safety
considered in design. suggests a tendency for sudden failure when the in-situ
stresses are diminished for any reason and is a warning
sign that care must be taken in terms of the excavation
and support installation sequence. For this reason, many
tunnel designers consider that it is prudent to design the
tunnel support on the basis that there are no in situ
el/zo/F/a
Figure 5. Example of the effect of in-situ stresses (40 m deep and kg = 1) on a large unstable wedge in a large span and
unsupported cavern.
stresses ensuring that, for almost all cases, the support 3.2.1 The design myth
design will be conservative.
There is significant misunderstanding and design myths
Although some level of conservatism is generally on how rock bolts provide such reinforcement effects. To
desirable, one main difference should be considered in this end, Pells (2008) stated that rock bolts are sometimes
the support of large span tunnels and caverns. In almost ascribed abilities that verge on magic. For example, they
all cases, these large span excavations will be are said to prevent stress induced failure, or said to
sequentially excavated such that large deformation that interlock a rock mass like aggregate. The latter is based
could induce wedge dislodgment would typically be on Tom Lang’s famous 1960‘s upside-down bucket
controlled and some anchorage and shear reinforcement experiment during construction of the Snowy Mountain
would already be in place prior to full exposure of the project which initiated a design myth that is often misused.
largest possible wedge, making the use of the positive Tom Lang‘s experiment intended to demonstrate to the
clamping effect of stresses more amenable. Figure 6 Snowy workers how rock bolts work.
depicts this scenario.
Evert Hoek visited the Snowy project in the 1960’s and he
An example of such large span with considerably shorter was impressed with such a demonstration (Hoek, 2007).
bolts in comparison to the excavation span is the Gjevik He then started using his own version for teaching
Olympic Hall in Norway with a record-breaking span in purposes at the University of Toronto (Hoek, 2007). In
civil tunnels of approximately 61 m which only used a Hoek’s version, he indicated that a zone of compression
maximum bolt length of 12 m (Broch et al, 1996). is induced in the region shown in red in Figure 8 and this
provides effective reinforcement to the rock mass when
3.2 Reinforcement effect the rock bolt spacing, s, is less than 3 times the average
rock piece diameter, a, and the length, L, approximately
Another important factor is the change in focus from an 28.
anchorage or suspension effect of rock bolts to the
reinforcement effect within the tunnel roof. This concept of a compression zone promoted by rock
bolts was then extended to tunnels as a rock bolt
compression ring concept around the opening by Pender
et al (1963) as depicted in Figure 9.
Figure 8. Hoek’s educational version of the upside-down bucket experiment (after Hoek, 2007).
The above concept could be said relatively intuitive and It should be noted that the iterative design approach
simple to explain how rock bolts work. However, Pells proposed Oliveira and Paramaguru (2016) does not
(2008) demonstrated that Tom Lang's bucket experiment define the rock beam thickness and consequently rock
and consequently Hoek’s gravel table have little bolt length based on deflections defined as acceptable
relationship to the action of rock bolts around a tunnel prior to the analysis. In their approach, deflections are an
because the stress scale is all wrong. In these output, not an input, with the primary objective of the rock
experiments the confinement or compression provided by bolting design approach proposed by the authors being
the bolts is of the order of a few kilopascals which is too the development of the compressive arch within the
small, by several orders of magnitude, to have any effect “stitched” pseudo-equivalent rock beam. This is achieved
on the rock mass strength by “confinement” in a tunnel by assessing the excess shear forces developed within
scale. In addition, although Figure 9 has been reproduced the equivalent beam and comparing against the mobilised
in many text books, Pells (2008) also demonstrated via rock bolt forces estimated through methods such as those
analysis that similar scale issue is observed with this presented by Pells (2002). This approach has in generally
concept as shown in Figure 10. allowed for a reduction of bolt lengths in recent projects
when comparing to precedent designs as the effect of bolt
spacing is taken into
r"!
rum“
"‘
mam-mm
mmnhmngcnm
“'Y
rma
rd
r‘ 1
new 1 run)
’w
= ‘rm-MAA 0»... 1mm“. w-Iw um. ”A.”
r" T"
Figure 10 Contours of major and minor principal stress induced by 2m rock bolts at 1.1m centres pretensioned to 80 kN
(after Pe/ls, 2008).
r‘fi
Likely compression arch that provides stability
Zones where tension occurs such that the plates
W1
average rock piece diameter) rockbolt spacing
(mesh/shotcrete) have an important function.
rm
l);
”@fimrfire
At this scale, the double
plates/washers do help providing
J
rockbolt L
minor confinement/interlocking (red
length
I.
zone).
zone of compression
Figure 11 Likely compression arching that provides stability in Hoek’s experiment (modified after Hoek, 2007).
("w
h
pr“,
Figure 12 Compression arching that provides stability in laminated rock beams. Note the effect of the rock bolts by the red
T‘,
stress tensors, i.e. tensile stresses (after Ollveira and Paramaguru, 2016).
account and may be used to compensate for shorter bolt Figure 15 presents both ground reaction curves together
lengths. with the countors of total displacements around the
tunnel. The unsupported case shows that collpase
Arched profiles intitiates when the excavation boundary tractions are
equivalent to approximately 15% of the original in-situ
In the case of arched profiles, the objective of the rock stress which occurs approximately 5 in behind the
reinforcement is no different, i.e. the focus remains on excavation face. At this point the crown deflection is
facilitating the stress redistribution around the tunnel. appproximately 70 mm. Full collapse is observed at a
boundary traction equivalent to 10% of the original in—situ
A discontinuum model of a 20 m span shallow excavation stresses which occurs approximately 7m behind the
in fair to poor quality siltstone rock (UCS=4-8MPa and excavation face. On the other hand, the rock bolted case
GSI = 40) using 3DEC will be adopted to illustrate the indicates that at a boundary traction equivalent to
effect of rock bolt reinforcement (Figure 14). Typical approximately 15% of the original in-situ stress or 5 m
parameters to those assumed in Sydney are used for the behind the excavation face, the crown displacement is
purpose ofthe analysis with two cases analysed: (1) afull- less than 35 mm. The tunnel reaches full stability some
face unsupported excavation and (2) a full-face 10 m behind the excavation face with a crown
excavation supported by rock bolts only spaced at 1.25 m displacement of approximately 75 mm when the
centre to centre. The Ground Reaction Curve of the two boundary traction is equivalent to 5% of the original in-situ
cases are assessed by simulating a gradual excavation stress. After this stage, the boundary traction is reduced
which is achieved through the reduction of the surface to zero in the rock bolted case simulating full excavation
tractions at the boundary of the excavation. and a distance greater than 60 m behind the excavation
Figure 13. Reinforcement effect for equivalent rock bolt stitched rock beam (after Oliveira and Paramaguru, 2016).
Geometry
Bolted rock arch — 4m bolts
Figure 14. Model of a 20 m span arched shallow tunnel in in fair to poor quality siltstone rock.
face with no further significant deflections observed in the to approximately 1 MPa. In fact, the zones with 1 MPa
tunnel crown. stresses indicate the likely line of thrust associated with
stress arching.
It should be noted that the large deflections observed in
the rock bolted case of Figure 15 would be futher reduced An analogy between conventional structural tunnel
if additional passive support such as a thick layer of support and the overall effect of rock reinforcement could
shotcrete (e.g. t> 300 mm) is applied which is generally then be made by treating the reinforced rock beam or
the case in shallow tunnels. Shotcrete would particularly reinforced arch as structural elements where loads are
control the fall-out of wedges between bolts and reduce applied with appropriate factors to test compliance to
overal rock loosening. design standards.
Based on the discussion above, the positive effect of the For example, using the previous example, a load
rock reinforcement in promoting a more stable stress equivalent to approximately 15% of the original in-situ
redistribution is evident through the comparison between stress, i.e. equivalent to the collapse initiation load, could
the ground reaction curves and equivalent deflections at be assumed and multiplied by a load factor of 1.5 in
several points behind the excavation face. accordance with AS5100. Such loads could then be
applied to a bolted rock arch modelled isolated from the
The improvement in stress arching promoted by the bolts rest of rock mass to verify its ability to promote the stress
becomes even more evident when comparing the stress arching and form the line of thrust, thus, providing support
tensor within the bolted-arch for the two cases. This is to the overlying rock mass as illustrated in Figure 17. In
illustrated in Figure 16 with the stress tensor coloured with this case, the structural capacity of the rock bolts would
respect to the magnitude of the major principal stress also have a reduction factor applied to satisfy code
(Minimum Principal in 3DEC as compression is negative). compliance.
The major principal stress indicates the ability of the rock
to redistribute or arch the stresses around the tunnel. For It is important to note, that similarly to the approach
the unsupported case, a large zone in the immediate roof adopted for rock beam design (Oliveira and Paramaguru,
of the tunnel is observed to be in tension or at zero 2016), the bolted-arch should first be allowed to deflect
stresses at a boundary traction equivalent to 10% of the elastically while maintaining a non-zero tensile strength
original in~situ stresses which confirms the loss of support within the joints. This initial elastic deflection is assumed
and associated collapse. On the other hand, the rock equivalent to the gradual excavation mechanism so that
bolted case indicates that the major principal stress in the some deformation occur before the roof is fully excavated
immediate roof of the tunnel is approximately 600 kPa at and formed.
a similar 10% boundary traction stage, locally increasing
fa
rm
Contour 0f Displacement
,9...
Deformed Factor: 5
100 8.0000502
7,0000E-02
Percentage of in-situ stress (96)
90 Unsupported 6.0000602
rww
5.0000502
80 Rockbolts
4.0000E-02
70 3,0000E—02
7'"
2.0000E—02
60 1.0000E—02
50 0.0000E+00
W‘
40
3O
r1
20
10
rs
@fifir‘r‘roflm
Stress
Scale: 19-06
100 Minimum Prin.
0.0000E+00
Percentage of in—situ stress (%)
90 —-—-Unsuppotted -1.0000E+05
-2.0000E+05
80 Rocklnlts 6.0000905
Unsupported: 4.0000E+05
70
~5.OOOOE'I05
observe section —6.0000E+05
60
going in tension 10000905
50 3.0000905
8.0000605
40 -1.0000E*06
rm:
30
20
10
r1I
,
\
\
\
"1
bolts in facilitating stress redistribution
but not as compression bulbs around The line of compression/thrust can be
the bolts... seen through the 1 MPa values...
Figure 16. Ground reaction curve and representative behaviour with respect to stress tensor within “bolted-arch”.
After the initial elastic equilibrium is achieved, the joint span/4 for spans greater than 18 m as proposed by Lang
tensile strength should be set to zero and the arch allowed (1958) and the curves proposed by Barton et al (1974)
to continue deforming until either equilibrium or failure and Barton et al (1995).
occurs.
For example, long tunnels with spans of up to 20 m have
4 UPDATED EMPIRICAL CHART been recently designed with bolt lengths of 5 m in contrast
to the precedent design of 6 m long bolts. Although this
Figure 18 presents an updated empirical design chart may seem a small reduction at first, in a long tunnel, this
including recent projects where some of the concepts results in savings of several kilometres of rock bolts which
discussed above have been applied. There is an evident include drilling and grouting and the associated time
increase in points below the previous precedent design required to install a large number of them.
trends and approaching the low bound curves equal to
‘‘ »
Cnmpresion arch/thrust transferreg,” %{ force and bedding
through bedding ~ e ‘4
10
8 .
: x
7 ., c
3 y = 0.5983x0766‘
g e
v
R2 = 0.875 . /
III
_: PRECEDENTED DESIGN AREA ,.o
... ., A 01
E’ 5
a)
a 2 O
. ''
-I x .x
g
o 4
03 A A
3 r " ——l
/ ,
1% 1" UNPRECEDEN TED DESIGN AREA
2 w- 7! x
1
O 5 10 15 20 25 30
Span (m)
Span/4 (Lang 1958) 2 * 015 X Span (Barton at at 1976) -—-1 4 #0184 x Span (Barton etal 1995)
0 Existing M5 9 80nd: Pump Station )1 Opera Hourse Car Park
2: M2 Motorway A Eastam Drstnbutor o Elgas Drstnbutor
1- Nonnsrde Storage Tunnel - Poatma A Lane Cove Tunnel
u Energy Austraua Cable Tunnel A ECRL x Transgnd Cable Tunnel
x Transgnd Plant Room 0 Buranda Street 4» 80990 Rd
- Coal Mines - NSBT o Brunswrck Street
NST APL Cross City‘l'unnel‘
New Southem Ralway' I Lutwycne Cavern :: Kodron Cavern
o WestConnex Stage 2 A WestConnex Stage 18 o NorthConnex
- . . - . Line of Best Fit Span/3 — Linear arch (E=l SGPa/q=50kPaJden <20mm)
6 ACKNOWLEDGEMENTS
REFERENCES
ABSTRACT
The design and construction of semi flat-roofed tunnels, i.e. with a high arch radius to roof span ratio, using a voussoir
beam analogy has been proven successful over time. In spite of such a success, the linear arch theory or voussoir beam
analogy has always been subjected to a certain level of scepticism due to some of its perceived limitations. Some of the
concerns are related to appropriate design methods for the design of rock bolting of multiple beds/laminations in cases
where single laminations are deemed unstable upon excavation or while addressing some adverse conditions. This paper
investigates the applicability of an analytical solution of the voussoir beam theory for the design of rock bolts in
laminated rock beams which has been confirmed with numerical analysis using DEM (Distinct Element Method)
analysis. The proposed analysis method can be easily implemented in a spreadsheet to provide rapid assessments
though it is considered only one part of the design process with other potential instability mechanisms assessed using
other analysis methods.
1 INTRODUCTION
Despite the successful design and construction of semi flat-roofed tunnels over the years in Australia, Canada and the
US, the linear arch theory or voussoir beam analogy has always been subjected to a certain level of scepticism due to
some of its perceived limitations. According to Diederich and Kaiser (1999), it generated a great deal of controversy
when it was first published by Evans (1941), even though the general notion of its successful application traces back to
ancient Rome architecture.
A significant portion of the scepticism seems to be related to perceived limitations of the conceptual model and
available analytical solutions with respect to:
§ The effect of horizontal stress
§ Span to lamination/bed thickness ratio
§ The presence of adverse geological features
§ The effect of a slightly arched roof
§ Bolting of multiple beds/laminations in cases where single laminations are unstable upon excavation
It is important to note upfront that there are two primary factors that promote a geological environment amenable to a
voussoir beam analogy in tunnel support design: (1) a horizontally bedded rock mass with no low to mid angle jointing
cross-cutting the lamination, i.e. a reasonably good quality rock mass, and (2) the existence of favourable horizontal
stresses.
This paper investigates the applicability of an analytical solution of the voussoir beam theory through comparison with
numerical modelling that focuses on the above limitations. The results illustrate that the voussoir beam analogy can be
confidently used in practice when used with reasonable engineering judgment as required in any other design method. It
adequately represents the results of more robust numerical solutions such as Discrete Element Method (DEM) when
similar governing mechanisms are considered in both models.
Line of thrust
a distance of approximately !/(2 2) from the midspan where s is the total span of the beam. This point is located
where the line of thrust crosses the centreline of the beam and it is reasonable to expect that the entire beam section is
under compression and that this stress is constant across the entire beam thickness T. As a result, the stress distribution
along the line of thrust can be represented by a two-part parabolic curve given by:
Line of thrust
Figure 2: Parabolic compressive stress variation assumed by Diederichs and Kaiser (1999) against Brady and Brown
(1993)
where x is the distance along the rock beam and the quadratic parameters may be found from the boundary conditions
given in Figure 2 resulting in:
where fm is the maximum compressive stress is given by Diederichs and Kaiser (1999). The geometric position of the
line of thrust is given by:
where T is the lamination thickness and N is the depth ratio for the triangular stress block according to Diederichs and
Kaiser (1999). These two equations will play a significant role in the assessment of multiple laminations although not
directly and explicitly used in the analysis of a single lamination.
sj T
Booker and Best (1990) developed a 1D finite element method which accounted for the effect of horizontal stresses
through a “cracked beam” analysis. This method was used for the design of the Sydney Opera House Carpark (Pells et
al., 1994) and seemed to capture the end effects in a more realistic manner. However, this method involves a finite
element solution that is inherently more sophisticated and less used than the Diederichs and Kaiser (1999) solution. As
a result, it will not be used or discussed in this paper.
where scb is the block compressive strength at the appropriate scale, a is an empirical exponent to account for the
confinement effect, sh is the initial horizontal stress and Esb is the horizontal rock mass Young’s modulus of a single
bed or lamination which is the relevant modulus for the analytical solution. This value may be estimated by:
where Eib is the modulus of the “intact” rock block at the appropriate scale, kn is the normal stiffness of the subvertical
joints at a corresponding spacing sj. In estimating Esb,modified, an upper bound value of 1.5 to 2.5Eib is proposed for
Equation (3). For most of the analyses in this paper, a limiting value of 1.5Eib will be adopted.
Based on the database used by Ramamurthy (1995), the exponent a could be said to vary between 0.06 and 0.2 (Figure
4). However, for the effect on a voussoir beam, where the block strength to confining stress ratio, i.e. scb/sh, is typically
low, a value between 0.03 and 0.06 seems to provide reasonable results compared to numerical models as demonstrated
as follows. For the low end values of a = 0.03, a lower cut-off value of 1.5Eib is recommended whereas 2.5Eib would be
recommended for the high end value of a = 0.06.
Esb/Esb,modified
a=
sci/sh
Figure 4: Confinement effect on rock mass modulus (modified from Ramamurthy, 1995).
s
h
s
h
s
h
s
h
s
h
s
h
s
h
s
h
s
h
s
h
Figure 6: Maximum beam deflection versus horizontal stress for various span-to-bed thickness ratios.
It is important to note that the DEM models require small element size definition to better capture the stress arching and
appropriate deflection prediction as pointed out by Oliveira and Pells (2014).
Figure 8: Examples of failure mechanisms affecting the immediate tunnel roof rock beam.
Oliveira and Pells (2014) also highlighted that it is unreasonable to only contemplate design scenarios that are
favourable for the voussoir beam analogy and require “perfect knowledge” of a “perfect rock mass”. The analyses must
give cognisance to the fundamental importance of rock bolting in providing a robust design against the uncertainties in
joint directions that may be expected in strata such as the Hawkesbury Sandstone of Sydney, the Bunter Sandstone of
the UK, the Beaufort Series in South Africa and the sandstones at Poatina in Tasmania (Oliveira and Pells, 2014).
Figure 8 presents a few examples of failure mechanisms that need to be addressed as part of the rock bolting design
process. Figure 8a shows a classical problem of potentially unstable rock blocks formed by an arched shape in bedded
rock. This particular example will be further investigated later in this paper with particular attention to the applicability
of the voussoir beam analogy. Mechanisms (a) to (d) of Figure 8 illustrate potentially unstable areas (shaded in red) that
would rely on a better performance of upper rock beams. In such cases, the unstable blocks and slabs maybe considered
as dead loads for the upper rock beams as a preliminary analysis approach. Figure 8e illustrates a more conventional
issue of the voussoir beam analogy for cases where thin laminations would be unstable at a certain span when behaving
as single rock beams, therefore requiring to be bolted to the upper roof.
partings as illustrated in Figure 9, therefore, controlling shear and slip along the bedding partings that would otherwise
result in delamination.
Figure 9: Equivalent rock bolt stitched rock beam (sub-vertical joints not shown for clarity).
The equivalent voussoir beam compressive stresses, f, may be estimated using the same methodology described for a
single bed with the difference that the new equivalent beam thickness, Teq, corresponds to the sum of the multiple
laminations. The development of the compression arch within the equivalent beam will generate both normal and shear
forces along the bedding partings which can be estimated by the relationships suggested by Asche and Lechner (2003)
given by: Another important aspect in the rock bolt design process is to consider that the mobilisation of the majority of
the reinforcement effect generally occurs upon some rock beam deflection and associated shear displacement along the
beddings despite some bolt pretension. As a result, in order to assess the mobilised bolt forces, it is necessary to
estimate the shear displacements resulting from the shear strain induced by the compression arch within the equivalent
beam. The shear displacements are then given by:
where
f(x) is the compressive stresses estimated by Equation (1), Bt is a normal stress induced by any rock bolt pretension
before significant rock beam deflections and q(x) is the angle between the line of thrust and the horizontal bedding
parting which may be found by differentiating Equation (2) resulting in:
The available bedding parting shear resistance and consequently excess shear can now be estimated using Equations (5)
to (7). The excess shear stress is then used to assess the required rock bolting forces to “trick” the multiple laminations
into behaving as an equivalent thicker rock beam and is given by:
where fb and cb are the friction angle and apparent cohesion of the bedding partings.
An important factor in the assessment of the equivalent bolted voussoir beam behaviour is a reasonable estimate of the
effect of the bedding partings on the overall rock beam deflections. The presence of the bedding partings and associated
shear displacements upon deflection mean that the rock mass modulus can no longer be estimated by Equation (4)
which only accounts for the normal stiffness of the cross-cutting joints. This estimate could be done by a direct
assumption of a modulus ratio, Esb/Gbeq, which often varies between 10 and 20 for transversely isotropic rock masses.
However, the authors of this paper have found that the following equation provides a reasonable estimate of the
equivalent voussoir beam modulus, Ebeq:
where ks is the shear stiffness of the bedding partings, n is the rock mass Poisson’s ratio (typically n = 0.25), Tsb is the
average thickness of the individual laminations or bedding parting spacing and all other parameters as previously
defined.
Another important aspect in the rock bolt design process is to consider that the mobilisation of the majority of the
reinforcement effect generally occurs upon some rock beam deflection and associated shear displacement along the
beddings despite some bolt pretension. As a result, in order to assess the mobilised bolt forces, it is necessary to
estimate the shear displacements resulting from the shear strain induced by the compression arch within the equivalent
beam. The shear displacements are then given by:
It is important to note that during this design process, an upper bound for the deflection of the rock bolted laminated
beam is the deflection of a single lamination. In other words, in general, multiple laminations stitched by rock bolts
cannot deflect more than a single bed as this would mean delamination of the beams.
3m
Figure 11: Analytical results for 3 m thick beam with sh = 0 MPa, bolts at 90º and ks = 10000 MPa/m.
As an attempt to improve the performance of the stitched laminated beam, the bolts are now considered to be installed
at a 70º angle to horizontal as shown in Figure 12. At such an angle, the rock bolts are likely to intersect the sub-vertical
joints and consequently generate additional axial forces. The bolts may also intersect multiple beddings within the
voussoir compression arch length at slightly different locations (Figure 17). For simplicity, to account for the bolts
crossing multiple beds, the total mobilised bolt force can be multiplied by MAX[N.(nb-1),1] where N is the depth ratio
for the triangular stress block and nb the number of beds. In addition, assuming that the rock bolts intersect the sub-
vertical joint near the bolt-bedding plane intersection, it can be considered reasonable to assume that this additional
axial force is similar to that developed within a bedding plane. This has been confirmed in the numerical models (Figure
17). The latter additional axial forces can then be converted to additional normal forces and consequently frictional
resistance.
The results are provided in Figure 13. The rock beam behaviour remains essentially the same as the stiffness provided
by the rock bolts added to that of the bedding has only a small influence on the beam deflections. However, the rate of
rock bolt forces mobilisation is significantly higher which results in more rock bolt forces mobilised for the same shear
displacement. For example, full bolt shear resistance is mobilised at approximately 1.2 mm. In addition, due to the rock
bolts intersecting the sub-vertical joints, additional axial forces were included in the shear resistance as discussed above.
However, the mobilised bolt shear forces/stresses are below the excess bedding shear stress which means that the
compressive arch in the equivalent beam cannot develop.
Figure 13: Analytical results for 3 m thick beam with sh = 0 MPa, bolts at 70º and ks = 10000 MPa/m.
One could ask what would be the consequences of the compressive arch not being formed in the previous analysis. As
depicted in Figure 9, the compressive arch generates both shear and normal forces across on the bedding parting. If such
forces are not resisted by the combined bedding and rock bolting shear resistance, plastic shear displacement or slippage
along the bedding parting takes place as illustrated by the red arrows in Figure 14. Such plastic shear displacements
induce further deflection of the rock beam which in turn could mobilise additional rock bolt shear resistance. A
simplified approach to capture such a mechanism is to artificially and iteratively reduce the bedding parting shear
stiffness which in turn reduces the equivalent thicker voussoir beam modulus (Equation 9). This process, as illustrated
in Figure 14, is carried until there is equilibrium between the excess shear forces and the rock bolt mobilised forces or
until the beam behaves as single laminations, if stable, or become unstable.
i+1 i+2
ksi+1
ksi
ksi+2
us
Figure 14: Proposed iterative approach for rock bolt mobilisation upon plastic shear displacements.
Figure 15 presents the results of the previous analysis adopting the proposed iterative approach. Equilibrium between
excess shear force and rock bolt mobilised shear forces is reached when the bedding shear stiffness is reduced to
approximately ks = 1500 MPa/m with a corresponding beam deflection of approximately 11 mm. The rock bolt
mobilisation curve is the same is that presented in Figure 13.
Figure 15: Analytical results for 3 m thick beam with sh = 0 MPa, bolts at 70º and ks = 1500 MPa/m.
A comparison between the analytical approach proposed above and DEM models is presented in Figure 16. A single
bed analysis is presented for reference. Like the analytical solution, the DEM model also indicates that the bolts do not
mobilise enough shear resistance for the development of the thicker rock beam if installed at a 90º angle. The beam
behaviour is compatible with that predicted by the analytical approach and similar to that of a single lamination which
is the limiting deflection for a stable beam.
Figure 17 shows the stress arching in the bolt stitched 3 m beam with rock bolts installed at 70º, sh = 0 MPa and
ks = 10000 MPa/m. The resulting deflection is approximately 11 mm, similar to the analytical prediction. The
reinforcement effect of the rock bolts and tensile zones (or zero stress) developed in the beam are evident through the
stress tensors in red colour, and consistent with the conceptual model discussed above. The effect of the rock bolts on
the sub-vertical joints can also be observed.
Figure 16: Analytical versus UDEC models for 3 m thick beam with bolts at 90º and 70º.
Figure 17: Stress tensor illustrating arching in DEM model - 3 m thick beam, bolts at 70º, sh = 0 MPa and
ks = 10000 MPa/m.
In summary, both analytical and UDEC models show similar trends in all cases analysed despite the small differences
in deflection (1-4 mm) likely resulting from the simplified assumptions on the effect of initial horizontal stresses,
equivalent beam modulus and also the differences in how the rock bolts are modelled in both solutions. As a result, the
analytical approach is considered to provide a reasonably satisfactory prediction of the overall performance of the
bolted laminated rock beam.
Figure 18: Behaviour of multiple beams with fixed end (after Obert and Duvall, 1967).
Based on a number of assumptions but mainly that the beams would have the same deflection over their entire length,
Obert and Duvall (1967) suggested that for a simplified two beam system the additional load to be added on the lower
beam and subtracted from the upper beam, could be estimated by:
where the subscripts “u” and “l” refer to the upper and lower beams respectively, w is the beam self-weight as a
uniformly distributed load, I is the second moment of area of the beam and E the deformation modulus. Similar
estimates could be made for an assumed number of beds. However, Obert and Duvall (1967) have not considered the
effect of sub-vertical cross-cutting joints and how it would affect the values of “l”, i.e. the voussoir analogy. In addition,
the designer would still have to assume a certain number of beams in the roof.
The authors of this paper have found that, as a first guess, a parabolic surcharge equivalent to an overburden of
approximately 0.25 times the span of the beam provides a reasonable and safe first estimate when comparing to DEM
models with enough rock cover. It is important to note that upon increasing deflections, one could consider that such a
surcharge reduces up to a deflection corresponding to full delamination where the single beds are assessed stable and
sustaining their own weight. If the material above the rock beam is considered unable to arch, this load would remain
constant upon deflection. In addition, this loading percentage could increase to 1 times the span above the rock beam
for a tunnel cover in soft ground, i.e. soil.
However it would still be necessary to assess the stress redistribution around the excavation and associated failure
mechanisms suchs as lamination buckling when behaving as column (Euler buckling), spalling and others. In doing so,
an equivalent elastic continuum analysis could be carried out, particularly considering a transversely isotropic
behaviour where elastic anisotropy related to the differences in vertical and horizontal modulus and a ratio of E/G = 10
to 20 can be captured. This can be easily achieved using programs such as Examine2D from Rocscience. The results of
this equivalent continuum analysis can then be used as an alternative check. The designer may verify the total shear
stress induced by the tunnel excavation and the normal stresses acting on the bedding partings at a certain depth into the
roof, say 0.5 m to 1 m, therefore being able to estimate the excess shear stress. The rock bolts assessed with the current
voussoir beam analogy approach can then be verified against the excess shear stress developed in the equivalent
continuum analysis.
In doing both verifications, the designer will target two primary objectives: (1) satisfy the formation of the compressive
arch in the voussoir beam analogy, and (2) attempt to maintain the roof (laminated rock beam) behaviour as close to an
elastic behaviour as practically possible where deflections are controlled to acceptable levels.
contact with the rock beam (Figure 20). This is conversion is consistent with the solution of the voussoir beam targeting
a compensated moment generated at the abutment due to self-weight of the beam with a resisting moment as the beam
deflects (Diederichs and Kaiser, 1999).
Assuming that some material has been removed from the beam (Figure 19), Asche and Lechner (2003) suggested that
the lowest stress, used to calculate the average stress, must be higher than that presented by Diederichs and Kaiser
(1999) as depicted in Figure 2. As a result, assuming that the force is unchanged, the new lowest stress could be
estimated by;
It is important to note that the above corrections are made within the iterative approach proposed by Diederich and
Kaiser (1999). This means that the final value of fm found with the iterative approach is already the corrected value and
no further adjustments need to be made in Equation (1).
Figure 21 shows the maximum beam deflection against the horizontal stress for the arched tunnel roof analysis. The
differences in deflections may be the result of two main assumptions: (1) the simplified approach adopted to account for
the effect of the initial horizontal stress and (2) the estimates of rock beam deformation modulus are based on average
spacing of the discontinuities, i.e. bedding partings and joints, which are no longer constant in the case analysed. The
discrete discontinuities better captured in the DEM model result in stiffer rock beams and therefore with smaller
deflections. Nevertheless, the maximum deflections predicted are considered to show a satisfactory agreement with
similar trends. The analytical solution seems to over predict the beam deflections compared to that of the numerical
solution due to the factors discussed previously. It should be noted that one rock bolt per rock wedge have been
introduced at 90º angle to provide stability to the blocks underneath the beam and these have affected to the
deformation of the beam. The analytical solution is not sensitive to initial horizontal stresses greater than 0.5 MPa
whereas the numerical model still indicates some minor reduction in deflections.
A second case was analysed considering that the tunnel roof arch creates an undercut on the first “continuous” rock
beam in the roof, leaving a beam thickness of only 0.75 m as depicted in Figure 22. Similar to “Case 1”, as an initial
assumption, the potentially unstable wedges are considered as an equivalent distributed load applied to the upper rock
beam with a corresponding value of q = 4 kPa.
Figure 21: Maximum beam deflection for arched tunnel roof - case 1.
Figure 23: Maximum beam deflection for arched tunnel roof - case 2.
Again, the differences in deflections may be explained by the simplified approaches adopted for both horizontal stresses
and rock bolt mobilisation. The combined cut-off values of 1.5Eib and a = 0.03 previously adopted for the effect of
initial horizontal stresses have a particular effect which can be observed in the shape of the curve of Figure 21 and
Figure 23. To investigate this effect, additional analyses were carried out assuming an increased cut-off value of 2.5Eib
with a = 0.06. The results are shown in Figure 24. As expected, there is a better agreement with the DEM predictions,
indicating that for low stiffness rock beams such as those with low intact block modulus or with an undercut arch,
higher values of a and effect cut-off value may be more appropriate, i.e. a = 0.06 and 2.5Eib respectively.
Figure 24: Maximum beam deflection for arched tunnel roof with modified confinement factors- case 2.
8 CONCLUSIONS
A significant portion of the scepticism for the use of a voussoir beam analogy seems to be related to the perceived
limitations of the conceptual model and available analytical solutions with respect to the effect of horizontal stress, span
to lamination/bed thickness ratio, the presence of adverse geological features, the effect of a slightly arched roof. In
addition, there is also some concern on appropriate design methods for the design of rock bolting of multiple
beds/laminations in cases where single laminations are deemed unstable upon excavation.
The examples adopted in this paper sought to demonstrate that the effect of initial horizontal stresses may be taken into
account through a simplified approach though not strictly capturing the theoretical mechanism. They also demonstrated
that the solution still provides acceptable results for span to lamination thickness ratio less than 10, limit suggested by
Diederich and Kaiser (1999). The effect of an undercut arch was also investigated with satisfactory results found with
the analytical solution.
The primary objective of the rock bolting design approach adopted in this paper is to promote the development of the
compressive arch within the “stitched” equivalent rock beam. This is achieved by assessing the excess shear forces
developed within the equivalent beam and comparing with the mobilised rock bolt forces estimated through methods
such as those presented by Pells (2002).
As a result, this paper illustrates with several examples and discussion of the underlying assumptions that an analytical
solution of the voussoir beam analogy can be successfully used as a simplified tool for the assessment of roof support in
strongly bedded rocks where jointing orientations and horizontal stress conditions provide a favourable environment. It
is not suggested as the only means of assessment, particularly considering the significant advances in discontinuum
rock mass numerical modelling. However, the voussoir analogy provides the designers with a simple tool that can be
programmed into a spreadsheet and used to test scenarios and conditions in a quick manner, providing some insight into
expected behaviours.
9 REFERENCES
Asche, H.R. and Lechner, M.K. (2003). Design for the Cross City Tunnel, Sydney. Proceedings of the RETC. Seattle.
pp. 122-139.
Bertuzzi, R and Pells, P J N (2002). Design of rock bolt and shotcrete support of tunnel roofs in Sydney sandstone.
Australian Geomechanics, 37(3).
Booker and Best (1990). Coffey internal communication.
Carter, J. (2003). Pells analysis of the shear behaviour of a reinforced rock joint. Report by Advanced Geomechanics,
Sydney University,
Diederichs, M S and Kaiser P K. (1999): Stability Guidelines for Excavations in Laminated Ground - The Voussoir
Analogue Revisited, Int. J. Rock Mech. & Min. Sci.; 36, pp 97-118.
Evans W H. (1941). The strength of undermined strata. Trans. Inst. Min. Metall., 50:475-500.
Fayol, M. (1885). “Sur les movements de terrain provoques par l’exploitation des mines,” Bulletin de la Société de
l’industrie Minerale, 2nd Series, Vol. 14, pp. 818.
Itasca (2014). Universal Distinct Element Code User’s Guide.
Oliveira D. and Pells, P. (2014) Revisiting the applicability of voussoir beam theory for tunnel design in Sydney.
Australian Geomechanics Vol 49(3), pp 29-44, September
Pells, P J N, Best, R J, and Poulos, H G (1994). Design of roof support of the Sydney Opera House underground
parking station. Tunnelling and Underground Space Technology, Vol 9, No 2, pp201-207.
Pells, P. J. N. (2002). Developments in the design of tunnels and caverns in the Triassic rocks of the Sydney region.
International Journal of Rock Mechanics and Mining Sciences 39: 569-587
Peck, W A, Sainsbury, D P and Lee, M.F (2013). The importance of geology and roof shape on the stability of shallow
caverns. Australian Geomechanics, 48(3).
Ramamurthy, T. (1995). Strength and modulus response of anisotropic rocks. In Comprehensive rock engineering (ed.
J. A. Hudson), Oxford, Pergamon Press, Vol. 1, pp. 313–329.
a r t i c l e i n f o a b s t r a c t
Article history: Underlying much of Sydney, Australia is a composite rock formation known as the Hawkesbury
Received 19 December 2016 Sandstone. This unit is composed of clastic layers of variable competency including a number of thick
Accepted 9 January 2017 and strong layers. The presence of very high horizontal stresses within these layers is widely accepted.
Deep excavations such as basements or open cuts in Hawkesbury Sandstone often experience moderate
horizontal movements in excess of 1 mm per metre of rock excavation. These movements can result in
various scales of damage and excess loads on supporting elements. Stress induced failures in tunnels
and underground excavations have also been observed in a number of projects and include crushing,
spalling and or slabbing of intact rock blocks or shear failures associated with planes of weakness.
While most design approaches in ground engineering account for shear failure mechanisms, the assess-
ment of brittle failure is less common and less well understood. Conventional models and failure criteria
do not appropriately describe such behaviour and consequently the impacts on ground support may not
always be appropriately addressed. This paper presents some discussions on the modelling and assess-
ment of brittle failure in Hawkesbury Sandstone and some of the impacts of high in-situ stresses on tun-
nel support design.
Ó 2017 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.tust.2017.01.003
0886-7798/Ó 2017 Elsevier Ltd. All rights reserved.
D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23 11
consequently the impacts on ground support may not always be exhibit long term rock strength of approximately no more than half
appropriately addressed. of the UCS value at laboratory scale near excavation boundaries.
This paper presents some discussions on the modelling and With the above understanding, it becomes evident that stan-
assessment of brittle failure in Hawkesbury Sandstone and some dard shear failure criteria based on laboratory scale parameters
of the impacts of high in-situ stresses on tunnel support design. (e.g. conventional Mohr-Coulomb or Hoek-Brown) would over-
predict the field strength in relatively intact blocks as they do
not account for the strength reduction caused by the extensional
2. Brittle failure and conceptual models type damage initiation and propagation. As a result, a number of
studies have demonstrated that, for massive to blocky brittle rock
As discussed in Kaiser et al. (2010), various studies have demon- masses (GSI > 70), the rock blocks and rock mass strength is best
strated that brittle failure processes leading to breakouts often represented by either a tri-linear or s-shaped failure envelope as
dominate rock mass behaviour near excavation boundaries in high depicted in Fig. 2.
stress conditions rather than shear failure. In the context of rock According to Diederichs et al. (2004), Diederichs and Martin
failure, the term ‘‘brittle” is used to indicate spalling-type failure (2010), damage monitoring generally indicates that the low bound
due to tensile cracking or extensile fracture propagation leading field strength threshold is less sensitive to confining stress than the
to rapid cohesion loss, and not to describe the more general peak lab envelope. As a result, the limit for major principal stress
process of plastic post-peak strain-weakening and elastic, brittle- could be approximated by r1max = CI + (1–2)r3. This threshold
plastic shear failure. only applies at low confinements.
Diederichs (2014a,b) describes the mechanisms of transition At higher confining stress the strength envelope makes a tran-
from spalling to violent energy release known as rockbursting. sition up to the envelope defined by the yield or crack damage
Diederichs et al. (2010) also noted that spalling, while brittle in threshold (CD). This is based on the understanding that, at high
nature, need not be violent. In some cases, it could also be time confinements, micro-cracks initiated at grain-scale quickly
dependent. In unsupported conditions and under an anisotropic stabilize as they propagate away from the nucleation site, and
in situ stress field, the process of spalling can form notch geome- therefore the strength will be governed by crack coalescence and
tries often confused with wedge fallout. macro-scale shear failure.
At a laboratory scale, crack initiation (CI), as formally defined by The CD threshold is defined at the point of stress-axial strain
Diederichs and Martin (2010) is typically observed at approxi- non-linearity which coincides with the point of volumetric strain
mately 40–55% of the UCS, followed by a phase of stable crack reversal in uniaxial compression as shown in Fig. 1. The CI thresh-
growth and then crack coalescence and damage (CD) with old is generally observed at the point which a systematic increase
macro-scale shear failure (Fig. 1). However, sample heterogeneity in crack emissions follows an increase in applied stress and is also
increases at larger scales, thus, increasing local tensile failure and expressed by the point of lateral strain non-linearity (Fig. 1).
unstable cracking, i.e. coalescence. This explains why brittle rocks
Fig. 1. Typical stress versus strain plot for hard rocks in uniaxial compression showing interpretation of strain and acoustic emission counts related to damage evolution
(after Martin, 1997; Martin and Christiansson, 2009)
12 D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23
Fig. 3. Fracturing pattern after failure of virtual triaxial test samples (only compression shown).
Fig. 5. Potential spalling of a 0.5 m thick bed above crown with respect to the horizontal LDP and horizontal GRC curves.
c = 5 kPa, zero tensile strength, normal stiffness kn = 8 GPa and in Sydney. There is, however, some excavation distance between
shear stiffness ks = 0.8 GPa. The excavation was carried out in the onset of fracture initiation and well developed spalling failure.
advances of 1.5 m adopting a softening relaxation approach to sim- This is typical of spalling interaction with thinner laminations in
ulate the gradual mechanical excavation. The longitudinal dis- the roof.
placement profile was monitored on side walls and roof as well The above scenario indicates that spalling of beddings with
as the ground reaction curve. Negative values represent a distance thickness of 0.5 m or less are likely occur near the face while the
ahead of the face. No support was modelled in this example. roadheader is still operating. In fact, such a case has been observed
It is important to note that the initial condition and depth were in a number of projects such as the existing M5 tunnel. As depicted
selected such that the excavation induced stress redistribution in in Fig. 6, the roadheader pick trimming marks indicate that the
the roof would approach or slightly exceed that of the CI threshold. brittle failure of a 0.2 m thick bed in the roof occurred relatively
However, such condition can be considered relatively common for close to the excavation face.
tunnels that exceed 40–50 m depth in Sydney. In these cases, the tunnel support is installed after the spalling
Fig. 5 presents the model results as a function of the horizontal event. Some additional rock bolting outside of the pattern would
tunnel displacement ratio, i.e. the ratio between a certain displace- likely be required to support the ‘‘cantilevering” rock at the border
ment and the final total displacement, with respect to the excava- of the fractured volume. Although some significant movements
tion face (LDP). It also shows the corresponding horizontal stress along the bounding bedding parting would already have occurred,
relaxation, i.e. the ground reaction curve (GRC). monitoring for additional shear movements post support installa-
The yield indicators in the numerical model indicate that crack tion, for example using endoscopes, would be recommended. The
initiates at approximately 3 m from the excavation face and spal- durability performance of the rock bolts could be affected upon
ling likely to occur at approximately 4.5–6 m from the face. The significant shear movements along beddings. This will be further
corresponding horizontal stress relaxation factors were in excess discussed later in the paper.
of 60% and 80% respectively. Such distances are within a typical Despite the higher likelihood of failure of thinner beds close to
1-day excavation round for tunnels excavated with roadheaders the excavation face, a generalisation is not possible. Variable bed-
D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23 15
Fig. 6. Brittle failure observed during excavation of the existing M5 tunnel (photo
courtesy of Golder Associates).
Fig. 9. Potential spalling of a 1 m thick bed above crown with respect to the horizontal LDP and horizontal GRC curves.
discussed in Oliveira and Paramaguru (2016), but also have enough Pre-emptive design attempts to reduce some of the adverse
deformation capacity to withstand the sudden formation of the effects are not always practical due to the magnitude of the hori-
potentially dilational spalling zone. The bolts on the flans of the zontal stresses in Sydney and its unpredictability. A number of fac-
spalling zone must also have capacity for shear resistance and/or tors may affect both the occurrence and timing of the failure as
shear deformation as lateral convergence normally flows the cen- discussed above. Firstly, brittle failure cannot be avoided with rea-
tral failure of the laminations. The shotcrete also requires enough sonable practical measures. Secondly, the variability of in-situ
capacity to hold the fractured rock block that may be formed stress, quality and strength of the rock mass and presence of
between bolts. This requires sufficient energy absorption to sustain adverse features suggest that brittle failure could occur at different
some reasonable deflection. In addition, allowances for potential depths. Although it could be argued that its likelihood increases for
re-bolting and shotcrete repair would also be necessary. These con- depths in excess of 40–50 m, brittle failure of thinner bed may
siderations are further explored in the following section. occur and have been observed at shallower depths.
Nevertheless, some allowances can and have to be made as part
of the tunnel design and construction process to address some of
4. Impact on tunnel support design in Hawkesbury Sandstone the potential impacts. These can be divided into two main
categories:
The existence of relatively high horizontal stresses is often con-
sidered one of the key features for a geological environment amen- Safety issues
able to semi-flat roofed tunnels (Pells, 1993; Oliveira and Pells, Durability issues
2014). Based on a voussoir beam analogy, horizontal stresses typ-
ically have a positive ‘‘clamping” effect on the tunnel roof rock
beam, thus, controlling deflections and improving tunnel support 4.1. Safety concerns
(Oliveira and Paramaguru, 2016). However, adverse effects may
be expected if excessive shear displacements in the vicinity of As briefly discussed above, stress induced failures in the tunnel
the bolts and/or stress induced failure occur as discussed above. roof may result in large volumes of rock that need to be suspended
D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23 17
in the upper zone of the roof. In addition, some of these failures This case study illustrates the importance of considering poten-
may result in significant shear displacements that may overload tial stress-induced failure in assessing the volume of potentially
rock bolts. In the particular case of brittle failure, the shotcrete unstable rock that required to be suspended deeper into the roof.
design also needs to provide enough capacity to hold the fractured This volume also needs to be considered as a surcharge onto the
rock block that may develop between rock bolts. Some of these upper roof rock beam. Similar observation is valid for brittle fail-
aspects are discussed as follows. ures as shown later in the paper.
In addition to the impact of stress induced failures on the
anchorage mechanism of the rock bolts discussed above, another
relevant impact is on the reinforcement action mode of the bolts.
4.1.1. Rock bolts
It is known that rock bolts may be subjected to significant shear
The two ground failures observed in the Sydney LPG caverns
deformation along rock discontinuities before a structural failure.
(De Ambrosis and Kotze, 2004) provide useful insights for the
For example, Fig. 12 presents predictions of forces acting on a rock
design of the rock bolts subjected to stress induced failures. The
bolt installed at a 80° angle from the rock discontinuity plane and
tunnel was approximately 14 m wide and 11 m high and was being
subjected to shear movements. These predictions are based on the
excavated by drill and blast of a 5.5 m high top heading followed
closed-form solutions proposed by Pells (2002) and Carter (2003)
by removal of the bottom bench.
and assumed a solid bar diameter of 21.7 mm, modulus of
At the time of the first failure it was being excavated using a
200 GPa, yield strength of 600 MPa, axial rupture strain of 15%
pilot heading followed by removal of the side strips and being sup-
and a pretension of 50 kN. The bolt is assumed reinforcing a bed-
ported with 4.0 m long, hollow, high tensile steel rock bolts at a
ding plane with friction angle of 35° with zero dilation. It is
pattern of 2 m 2 m spacing (De Ambrosis and Kotze, 2004). The
observed that tensile rupture is estimated at shear displacements
bolts were initially anchored using a mechanical expanding shell
of approximately 38–47 mm.
at the top end of the bolts followed by the application of cement
Although not in sandstone, Li (2010) presented a few case stud-
grout applied through the hollow stem once the face had advanced
ies where rock bolts were subjected to high stress environments
and roof convergence had ceased.
and post-failure observation could be well documented. For exam-
The first failure occurred during or immediately after blasting of
ple, Fig. 13a illustrates a failed 20 mm rebar bolt subjected to a
a side trip when the pilot heading was some 5 m ahead of the side
shear movement of approximately 40 mm consistent with the pre-
strip removal. Considering that the rock bolts would only be
vious estimate, despite the smaller diameter.
grouted after convergence had ceased, the support was primarily
Another example is the Sydney LPG caverns (De Ambrosis and
relying on the mechanical anchors (i.e. as a friction bolt) to support
Kotze, 2004). After the first failure, the roof support was altered
wedge or rock block loads. However, the combination of blast-
to include the use of 5 m long, hollow, high tensile steel bolts fully
induced fractures and shear induced failure particularly controlled
grouted at the face on a denser grid spacing of 1.5 m (transverse)
by cross-bedding (potentially also blast affected), resulted in a
2.0 m (longitudinal). However, a second failure of similar magni-
large volume of unstable rock (Fig. 10). The thickness of the failed
tude occurred where a number of sheared rock bolt stubs were
sandstone slab was up to 1.5 m which overloaded the mechanical
observed at the failure surface in the crown of the cavern (Fig. 13b).
anchors unable to sustain such a load. As a result, a number of rock
Inspection of the failed rock on the floor of the tunnel showed that
bolts were simply pulled-out with the failed rock slab as depicted
the bolts had been fractured at the location of a bedding plane
in Fig. 11.
Fig. 10. Mapping of first roof collapse in the Sydney LPG caverns (modified after De Ambrosis and Kotze, 2004).
18 D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23
Fig. 13. Examples of structural rock bolt failure: (a) Metattliferous mine in Sweden due to combined tension-shear displacements (Li, 2010). (b) Sheared rock bolt at second
roof collapse in the Sydney LPG caverns (after De Ambrosis and Kotze, 2004).
Fig. 14. Shotcrete failure mechanism for trapezoidal fractured rock block.
failure is the governing mechanism, the limiting equivalent shot- required to be sustained by the shotcrete could be estimated using
crete confinement could be estimated according to Barrett and empirical evidence of bulking based on Kaiser et al., 1996 and
McCreath (1995) as: shown in Fig. 15. In this figure, the degree of bulking (volume
expansion during brittle failure) is a very sensitive function of con-
ra a
r3 ¼ 4 ð7Þ fining pressure (support pressure). Bulking factor is the average
s
percentage of volume increase within the yielded zone.
where ra is the shotcrete adhesive/bond stress, a is an assumed An initial estimate of the yielded zone may be carried out using
adhesive bond length and s is the rock bolt spacing. empirical relationships as that proposed by Diederichs and Martin
For example, the limiting shotcrete confinement is estimated at (2010) and depicted in Fig. 16.
approximately r3 = 7 kPa assuming a = 30 mm and ra = 0.1 MPa for
an early age shotcrete with a rock bolt spacing s = 1.75 m. Such
confinement is equivalent to an approximate load of 12 kN/m over
the rock bolt spacing. The maximum residual horizontal stress
associated with such a confinement is approximately r1 = 74 kPa
for a value of / = 56° for the fractured rock.
Assuming a reasonably conservative value of a = 30° with a typ-
ical maximum spalling depth of H = 0.5 m, the residual horizontal
stress provides a resisting force R = 57 kN/m. The block weight is
W = 31.5 kN/m for an assumed c = 24 kN/m3 which results in an
overall factor of safety FoS = 1.8. If the effect of confinement on
the internal stability of the wedge is ignored, the factor of safety
is estimated as FoS = r3/(W/s2) = 0.7, i.e. failure, which highlights
the positive effect of the very small confinement provided by the
shotcrete.
It is important to note that the above confinement effect relies
on the shotcrete being able to withstand the deflection imposed by
the effects of spalling and therefore with sufficient energy absorp- Fig. 15. Sensitivity to support pressure of (empirical) bulking factor as a result of
tion to maintain a certain residual strength. The likely deflections brittle failure.
20 D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23
For example, using a 2D approximation, if the average depth of guidance. In addition, the energy should also be corrected for shot-
damage/yield is 0.5 m above the roof, the estimated expansion into crete thickness and span (e.g. Bernard, 2013).
the tunnel would be 25 mm based on 5% bulking factor for bolts at This is consistent with experience in Sydney where lateral
1.75 1.75 m spacing (neglecting deformation tangential to or deformation upon spalling is not significant. It is worthwhile not-
parallel to the tunnel). This average expansion will be resisted by ing that such a deflection is associated with early ages of shotcrete
both the rock bolt and the shotcrete layer. While the total deforma- (i.e. <7 days) when it better performs upon large deformations
tion of the shotcrete layer will be more than this average, the rel- (Bernard, 2008). Nevertheless such a large deflection highlights
ative deflection between the bolt plates will be less. the need for repair.
Based on the restoring force required to be provided by the It is important to note however that other shotcrete failure
shotcrete discussed above, such a deflection would be equivalent mechanisms such as pure compressive failure and buckling also
to an energy absorption requirement of approximately 300 J at need to be investigated which is more appropriately carried out
the upper bound of 25 mm deflection (relative deflection of shot- through numerical modelling.
crete would be less than this, as discussed). This energy absorption
can generally be achieved with the use of fibre reinforcement (NGI,
4.1.3. Performance of the combined support
2013) alone without steel meshes. However, it should be
In order to investigate the above mechanisms, a case similar to
recognised that such energy requirement is not directly related
that analysed in the previous 3D analysis is re-analysed using a full
to the values achieved in test panels which therefore are used as
discontinuum model in UDEC.
Fig. 16. Empirical estimate of spalling depth (after Diederichs and Martin, 2010).
The Sandstone Class I/II is modelled with variable bedding strength, i.e. tensile strength, of 0.2 MPa, cohesion of 0.14 MPa,
thickness between 600 mm and 1.5 m and sub-vertical joint spac- friction of 35°, kn = 1 GPa and shear stiffness ks = 0.1 GPa.
ing varying from 2 to 4 m as depicted in Figs. 17 and 18. The rock After a relaxation of 95% the properties are increased to an
block was modelled assuming the CWFS parameters described in equivalent 7-day strength where E = 10 GPa, compressive strength
the previous 3D analysis. The bedding parting also follow the prop- of 24 MPa, tensile strength of 2.5 MPa and residual of 1.1 MPa. The
erties used in the 3D model whereas the joint deformation proper- shotcrete interface increases to bond strength of 0.3 MPa, cohesion
ties were adopted as kn = 10 GPa and shear stiffness ks = 1 GPa but of 0.21 MPa, friction of 35°, kn = 10 GPa and shear stiffness
with similar strength. ks = 1 GPa.
The tunnel is assumed to be excavated in full face with 3D The results are presented in Figs. 17 and 18. Brittle failure
effects simulated using a stress relaxation approach which controls occurs after installation of the tunnel support in agreement with
the excavation boundary forces. The tunnel support consists of the 3D model, considering that the immediate roof bed is
3.5 m long bolts at a spacing of 1.75 m by 1.75 m and 50 mm fibre approximately 0.5 m.
reinforced shotcrete which are installed after a stress relaxation of The effect of brittle failure in the overall behaviour of the tunnel
60%, i.e. stresses of approximately 40% of in-situ at time of support is reflected in the relatively large vertical movements which may
activation. be considered as a trigger level during monitoring of the excava-
The rock bolts are modelled using cable bolts in UDEC with the tion. The bulging, i.e. dilation, of the fractured rock and imposed
same parameters used to generate Fig. 12 including the allowance
of a rupture strain of 15% beyond which the capacity of the bolt
drops to zero. An ultimate bond stress of 2 MPa is assumed. The
dowel effect ignored in cable bolt elements is included with rein-
forcement elements adopting a shear stiffness, Ks, and shear capac-
ity, Ps,ult estimated by:
" #3=4
2Eg pffiffiffiffiffiffiffiffiffiffiffiffiffi
F bu rci
2
K s ¼ Eb I and Ps;ult ¼ 0:67d1 ð8Þ
4Eb Iðdd21 1Þ
where d1 is the rock bolt diameter (d1 = 21.7 mm), Gg is the grout
shear modulus (Gg = 12 GPa), Eb is the Young’s modulus of the rock
bolt (Eb = 200 GPa), d2 is the rock bolt hole diameter (d2 = 48 mm),
Eg is the grout Young’s modulus (Eg = 30 GPa), I is the rock bolt sec-
ond moment of area, Fby is the ultimate stress of the rock bolts
(Fbu = 920 MPa) and rci is the uniaxial compressive strength of the
intact rock (rci = 25 MPa). The axial effect of the reinforcement ele-
ments are set to zero.
Shotcrete properties were developed to account for the early
age. Between a stress relaxation of 60% and 95% it is modelled
adopting a Youngs modulus E = 6 GPa, Poison ratio t = 0.15, com-
pressive strength of 10 MPa, tensile strength of 1.2 MPa and resid-
ual of 1 MPa. The shotcrete interface is modelled assuming a bond Fig. 19. Benefits of split heading on tunnel support.
Fig. 20. Damage on cable bolt plastic sheathing after 15 mm and 18 mm direct shear movement (after Bertuzzi, 2004).
deflection on the shotcrete is also evident with the exaggerated cases. Large volumes of rock may be generated as a result of stress
displacements. The small confinement acting on the failure planes induced failure.
within the fractured rock is also confirmed although not shown When rock and rock mass fails in a brittle and more tensile
here. manner, the ground behaves different close to the excavation
The example confirms the expected performance of the adopted boundary than far from it. Rock fails by spalling rather than by
rock bolts where structural capacity is satisfied. However, the shot- shear. As a consequence, conventional modelling techniques and
crete and three of the central rock bolts experience significant constitutive models may be inadequate to design tunnels and their
shear displacement that may affect their long term durability as support systems, and may mislead designers leading to undesired
discussed in the following section. Due to the separation (i.e. construction delays and unexpected costs. The use of brittle mod-
deflection) of the immediate roof bed, the shear displacements els that account for cracking initiation and development, thus, dif-
within the central portion of the roof are no longer available for ferent mobilisation of cohesion and friction more appropriately
plotting. represent the failure mode.
It is important to note that a full face was modelled as a critical It has been demonstrated that the use of more appropriate
case. However, considering the large span there are some benefits models allows for a better understating of the support
in adopting a split heading excavation as illustrated in Fig. 19. requirements. However, the several triggering factors required
for the initiation of the brittle failure mean that where and when
4.2. Impact on durability it may occur is relatively unpredictable. As a result, the use of
generalised pre-emptive changes in the support may be onerous
Infrastructure tunnel projects typically require a minimum and unnecessary.
100 year design life for the permanent support which often A reasonable approach is to focus on safety concerns during
consists of rock bolts and shotcrete in rock tunnels. excavation with respect to the effects of stress induced failures
The typical approach for the design of rock bolts in Australia is such as large volumes of rock that need to be suspended and the
to address durability concerns by developing redudant corrosion bulking effect caused by spalling, both to be controlled with ade-
protection measures. It is generally accepted that a double corro- quate rock bolt spacing and shotcrete performance. The primary
sion protection system using plastic sheating as one of the protec- objective is to achieve a safe temporary excavation with construc-
tion provides the required durability. tion measures that allow repairs and reinforcements where and
As previously discussed, rock bolts may be subjected to signifi- when necessary upon an observational approach, particularly to
cant shear deformation along rock discontinuities before a struc- address durability concerns.
tural failure. On the other hand, direct shear tests on cable bolts
grouted within concrete blocks carried out by Bertuzzi (2004) indi- References
cated that damage on the plastic sheating protection may initiate
at shear displacements of less than 15 mm (Fig. 20). If such shear Barrett, S.V.L., McCreath, D.R., 1995. Shotcrete support design in blocky ground:
Towards a deterministic approach. Tunn. Undergr. Space Technol. 10 (1), 79–89.
displacment is observed along a bedding or joint that intersects a
Bernard, E.S., 2008. Embrittlement of Fiber Reinforced Shotcrete, Shotcrete, vol. 10,
rock bolt, the 100-year design life could be compromised. No. 3. American Shotcrete Association, pp. 16–21.
Similar issues are applicable to the shotcrete. Excessive move- Bernard, E.S., 2013. Development of a 1200-mm-diameter round panel test for post-
ment and deflection after application of the shotcrete could result crack assessment of fiber-reinforced concrete. Adv. Civil Eng. Mater. 2 (1), 457–
471.
in crack widths that lead to significant corrosion and reduction of Bertuzzi, R., 2004. 100-Year design life of rock bolts and shotcrete. In: Proceedings
the energy absorption capacity of the steel fibre reinforced of the 5th International Symposium on Ground Support, 28-30 September,
shotcrete. Perth, Western Australia.
Bertuzzi, R., 2014. Sydney sandstone and shale parameters for tunnel design. Aust.
As a result, allowance for both rock re-bolting and shotcrete Geomech. 49 (1).
repair and crack injection may be required in case excessive shear Bertuzzi, R., 2015. Forming a view of the rock mass strength of Hawkesbury
movements are observed during construction. Sandstone from tunnelling case histories. In: 13th International Congress of
Rock Mechanics, 10-13 May 2015, Montreal, Canada.
Brady, B.H.G., Brown, E.T., 2004. Rock Mechanics for Underground Mining. Springer,
5. Conclusions Dordrecht.
Carter, J., 2003. Pells Analysis of the Shear Behaviour of a Reinforced Rock Joint
Report by Advanced Geomechanics. Sydney University.
As demonstrated above, stress induced failures in Hawkesbury De Ambrosis, L.P., Kotze, G.P., 2004. Stress induced roof collapses during
Sandstone have been observed on a number of projects in Sydney. construction of the Sydney LPG storage cavern. In: Proc. 9th ANZ Conf. on
Such failures have an impact on tunnel support designs that need Geomechanics, Auckland, 8-11 February, pp. 159–165.
Diederichs, M.S., 1999. Instability of Hard Rockmasses: The Role of Tensile Damage
to be addressed. For instance, rock bolting is not only dependent on and Relaxation Ph.D. thesis. Dept. of Civil Engineering, University of Waterloo,
the existing block sizes as typically considered in gravity fall-out Waterloo, Canada.
D. Oliveira, M.S. Diederichs / Tunnelling and Underground Space Technology 64 (2017) 10–23 23
Diederichs, M.S., 2007. CGS Geocolloquium Award Lecture: Damage and spalling Martin, C.D., 1997. Seventeenth Canadian geotechnical colloquium: the effect of
prediction criteria for deep tunnelling. Can. Geotech. J. 44 (9), 1082–1116. cohesion loss and stress path on brittle rock strength. Can. Geotech. J. 34 (5),
Diederichs, M.S., Kaiser, P.K., Eberhardt, E., 2004. Damage initiation and propagation 698–725.
in hard rock tunnelling and the influence of near-face stress rotation. Int. J. Rock Martin, C.D., Christiansson, R., 2009. Estimating the potential for spalling around a
Mech. Mining Sci. 41 (5), 785–812. deep nuclear waste repository in crystalline rock. Int. J. Rock Mech. Min. Sci. 46
Diederichs, M.S., 2003. Rock fracture and collapse under low confinement (2), 219–228.
conditions. Rock Mech. Rock Eng. 36 (5), 339–381. McQueen, L.B., 2000. Stress relief effects in Sandstone in Sydney underground and
Diederichs, M.S., Carter, T., Martin, C.D., 2010. Practical rock spall prediction in deep excavations. In: McNally, G.H., Franklin, B.J. (Eds.), Proceedings of the
tunnel. In: Proceedings of World Tunnelling Congress ’10 – Vancouver, p. 8. ‘Sandstone City’ symposium at the 15th Australian Geological Convention.
Diederichs, M.S., Martin, C.D., 2010. Measurement of spalling parameters from Geological Society of Australia (Monograph No. 5).
laboratory testing. In: Proceedings of Eurock 2010, Lausanne, p. 4. McQueen, L.B., 2004. In situ rock stress and its effect in tunnels and deep
Diederichs, M.S., 2014a. Managing rockburst risk in D&B tunnels. In: Tunnelling excavations in Sydney. Aust. Geomech. 39 (3), 43–58.
Association of Canada Conference. Vancouver, p. 8. Oliveira, D.A.F., Wong, P.K., 2012. Selection of rock mass design parameters for
Diederichs, M.S. 2014b. When does brittle failure become violent? Spalling and assessing excavation induced movements in the Sydney CBD. In: Proceedings:
rockburst characterization for deep tunneling projects. In: Proceedings of the ANZ 2012 Conference, Melbourne, 15–18 July, pp. 789–795.
World Tunnel Congress 2014, Iguazu, Brazil, p. 10. Oliveira, D., Parker, C., 2014. An alternative approach for assessing in-situ stresses in
Enever, J.R, Walton, R.J., Windsor 1990. Stress regime in the Sydney basin and its Sydney. In: Proceedings: Australasian Tunnelling Conference 2014, Sydney, pp.
implications for excavation design and construction. Inst. Eng. Aust. VII 189–194. September.
Tunnelling Conf. Sydney, pp. 49–59. Oliveira, D., Pells, P., 2014. Revisiting the applicability of voussoir beam theory for
Enever, J.R., 1999. Near surface in situ stress and its counterpart at depth in the tunnel design in Sydney. Aust. Geomech. 49 (3), 29–44. September.
Sydney metropolitan area. In: Proc. 8th ANZ Conf. on Geomechanics, Hobart, pp. Oliveira, D., Chan, G., 2016. Ground control for a deep basement excavation in
591–597. Sydney’s GPO fault zone. In: Proceedings: Australian Geomechanics Society –
Hajiabdolmajid, V., Kaiser, P.K., Martin, C.D., 2002. Modelling brittle failure of rock. Sydney [Chapter 2016 Symposium, November].
Int. J. Rock Mech. Min. Sci. 39 (6), 731–741. Oliveira, D., Paramaguru, L., 2016. Laminated rock beam design for tunnel support.
Hewitt, P.H., McQueen, L.B., Davies, P.R., 1999. Genting Centre, Sydney – Deep Aust. Geomech. 51 (3).
Excavation Adjacent to Railway Tunnels. In: Proceedings of the 8th Australia NGI (Norwegian Geotechnical Institute) 2013. Rock Mass Classification and Support
New Zealand Conference on Geomechanics, Hobart,. Australian Geomechanics Design Handbook. NGI (<www.ngi.no>). p. 57.
Society, pp. 611–617. Pells, P.J.N., 1990. Stressed and displacements around deep basement in the Sydney
Kaiser, P.K., Amann, F., Steiner, W., 2010. How highly stressed brittle rock failure area. Seventh Australian Tunnelling Conference, Sydney, pp. 241–249.
impacts tunnel design. In: Zhao, J. et al. (Eds.), Proceedings of European rock Pells, P.J.N., 1993. Rock mechanics and engineering geology in the design of
mechanics symposium, Lausanne, pp. 27–38. underground works. EH Davis Mem. Lect., Aust. Geomech. 25. August.
Kaiser, P.K., McCreath, D., Tannant, D., 1996. Canadian Rockburst Support Pells, P.J.N., 2002. Developments in the design of tunnels and caverns in Triassic
Handbook. Geomechanics Research Centre and CMIRO Sudbury, Ontario. rocks of the Sydney region. Int. J. Rock Mech. Mining Sci. 39, 569–587.
Li, C.C., 2010. Field observations of rock bolts in high stress rock masses. Rock Mech. Walker, B., 2004. Stress relief on hillsides and hillside excavations. Aust. Geomech.
Rock Eng. 43, 491–496. 39 (3), 59–72.
REVISITING THE APPLICABILITY OF VOUSSOIR BEAM THEORY FOR
TUNNEL DESIGN IN SYDNEY
ABSTRACT
The design of semi flat-roofed tunnels, i.e. with a high arch radius to roof span ratio, in Sydney has been proven
successful over time. The horizontally bedded nature of Sydney’s Hawkesbury Sandstone draws designers to the
voussoir beam theory. Such analogy and the associated method of analysis can be easily implemented in computer
spreadsheets, which significantly facilitates the design of semi flat-roofed tunnels in geological conditions such as
Hawkesbury Sandstone. Like any other engineering simplified model or theory, the voussoir beam theory has some
limitations. However, it seems that some of these limitations are not well understood and often ignored and/or
misinterpreted. Such lack of understanding of the theory and its limitations often raises question about the applicability
of the analytical solution in practice, misleading engineers to believe that the only reliable and comprehensive design
method is through numerical analysis such as Distinct Element Method (DEM). This paper investigates the applicability
of the analytical or closed-form solution of the voussoir beam theory through comparison with numerical modelling,
focusing on some of the perceived limitations and their impact on the design of tunnels in Sydney. The results illustrate
that the voussoir beam theory can be confidently used in practice if its limitations are well understood and good
engineering judgment is applied to take the local geology into account. In addition, the results also demonstrate that
some of the limitations can be on the conservative side. For example, the potentially positive effect of high horizontal
stresses ignored in the voussoir beam theory may explain why some of the unfavourable conditions are less pronounced
in practice.
1 INTRODUCTION
According to Diederichs and Kaiser (1999), the notion of a linear arch theory often referred as voussoir beam was first
proposed by Evans (1941) specifically to explain the stability of a jointed or cracked beam. After some initial
controversy when first published, the voussoir beam analogue has been generally accepted and has since been revisited
by several authors (Pells et al., 1994; Sofianos, 1996; Diederichs and Kaiser, 1999) and presented as a simplified tool
for stability analysis of excavations in civil construction and in mining.
The horizontally bedded nature of the Hawkesbury Sandstone in Sydney draws designers to the voussoir beam or linear
arch theory. The theory describes the conditions for the stability of unsupported flat roof spans by means of a rock beam
analogy with joints sub-orthogonal to the tunnel span. It accounts for the arching effect of the vertical stresses and loads
upon deflection of the tunnel roof which in turn promotes compressive zones across the joints and consequently the
stability of the rock beam. Diederichs and Kaiser (1999) present the four main failure mechanism associated with a
voussoir beam as illustrated in Figure 1.
(c)
(a) (b)
(d) (e)
Figure 1: (a) Voussoir beam analogue with compression arch shown; Failure modes of the voussoir beam: (b) snap-
through; (c) crushing; (d) sliding and (e) diagonal cracking (modified after Diederichs and Kaiser, 1999).
The analytical solution is based on an iterative analysis of beam deflections which is then used for estimates of
maximum compressive stress, σmax, developed through the arching mechanism. Such estimates are then adopted to
assess factors of safety against crushing, sliding or shear at the abutments and a Buckling Limit (BL) Index for the
snap-through mechanism. Diederichs and Kaiser (1999) state that a BL index of 35% indicates yielding of the voussoir
beam and 80% collapse.
The voussoir beam theory has limitations like any other simplified engineering model or theory. However, it seems that
some of these limitations are not well understood and often ignored and/or misinterpreted which may raise questions
about its applicability in practice.
Figure 2: Sensitivity of the voussoir beam analogue to relatively minor geotechnical variations (after Peck et al., 2013
Figure 3: Comparison of the analytical solution adopted in this paper and Pells et al. (1994).
15 m
1m
γ = 26 kN/m3
Collapse
Yielding
Yielding
Collapse
1.4
Yielding
Collapse
1.2
1.0 1.0
1.0
0.8 0.8
0.8
0.6 0.6
0.6
0.4 0.4 0.4
0.2 0.2 0.2
0.0 0.0 0.0
0 50 100 150 0% 20% 40% 60% 80% 100% 0 5 10 15 20
Midspan
deflection
(mm) Buckling
limit
index
(BL) σmax (MPa)
ql 4 (2)
δ=
384EI
where q is the uniformly distributed load acting on the beam (in the base case only the self-weight), E is the relevant
beam Young’s modulus (i.e Eh,rm = 7.3 GPa) and I is the uncraked second moment of area of the beam.
STABLE
STABLE
(b)
(c)
Figure 6: Deflection contours for base case. (a) σh = 0 MPa; (b) σh = 4 MPa; (c) effect of initial horizontal stress.
It is important to note that the effect of horizontal stress has been modelled as an initial value, i.e. an in situ stress prior
to any excavation which changes only as a function of the rock beam deflection but not as a result of stress
redistribution of the excavation. In other words, it does not fully represent the stress redistribution and potential increase
induced by the tunnel excavation. However, the range of initial horizontal stresses adopted in the analyses is considered
representative of the expected horizontal stress values and roof beam behaviour for tunnels depths varying from 5 m to
40 m in Sydney’s Hawkesbury Sandstone.
STABLE
(a)
Deflection (mm)
0 4 8 12 16 20
STABLE
(b)
Figure 7: Deflection contours for base case showing finite difference zone sizes: (a) fine; (b) coarse.
Deflection (mm)
0 10 20 30 40 50
Figure 8: Deflection contours for single steeply dipping joint (dip 70º) with σh = 0 MPa.
A second case was run by the writers in UDEC considering a dip angle of 60º which could also be argued possible in
Hawkesbury Sandstone even though less frequent. Under such conditions, i.e. with an angle between the plane of the
cross-cutting discontinuity and the normal to discontinuities sub-parallel to the excavation plane equal to the friction
angle of the joints (30º), the UDEC model (Figure 9) predicts full collapse of the beams even at higher normal stresses.
However, it is important to recognize that due to current tunnelling standards in Sydney, it is normal practice to have a
minimum rock bolting pattern even in massive sandstone considered self-supporting. Such rock bolting pattern may
reduce such risk by reinforcing the single joint triggering the instability. The effect of rock bolts is investigated later in
this paper.
COLLAPSED
(a)
Deflection (mm)
0 10 20 30 40 50
COLLAPSED
(b)
Figure 9: Deflection contours for single steeply dipping joint (dip 60º): (a) σh = 0 MPa; (b) σh = 4 MPa.
Collapse
Yielding
Beam
thickness
(m)
Beam
thickness
(m)
Collapse
COLLAPSED
70
(a) Deflection (mm) 60
COLLAPSED 40
30 Collapse UDEC
(b)
20 Analytical
META-STABLE / YIELDING Elastic
beam
10
(c) 0
0 1 2 3 4 5
STABLE Horizontal
stress
(MPa)
(e)
(d)
Figure 11: Deflection contours for single non-persistent shallow angle joint. (a) σh = 0 MPa (b) σh = 1 MPa (c) σh =
1.75 MPa (d) σh = 4 MPa (e) effect of initial horizontal stress.
A relatively similar condition in Hawkesbury Sandstone could be argued to be the effect of cross-bedding partings on
composite beams, i.e. where low angle partings typically with dips of 15º to 30º (Bertuzzi and Pells, 2002a) exist above
a more massive bed and both are clamped together by rock bolts forming a composite thicker voussoir beam. However,
it is important to emphasize that this refers to the effect of cross-bedding partings, i.e. open cross beddings, and not
distinct intact cross bedding often observed in the fresher Hawkesbury Sandstone.
It is important to note that the effect of horizontal stresses is not necessarily always positive and high values of stress
may cause compressive buckling of thin beams, which may be assessed approximately by Euler’s buckling failure
mechanism given by:
π 2 Et 2 (3)
σ cr = 2
12lef
where σcr is the critical compressive buckling stress, E is the relevant beam Young’s modulus (i.e Eh,rm = 7.3 GPa), t is
the beam thickness, and lef is the effective length, equal to the length of the beam for pinned ends and half for fixed end
cases. Based on Equation (3), the critical buckling stress in this case would be between 6.7 MPa for a beam with pinned
ends and 26 MPa for fixed ends. As discussed above, it could be expected that the rock beam would behave more like a
fixed end beam under the effect of high horizontal stresses, therefore, tending towards the high bound value of the
buckling stress. However, it is important to consider that in this case, the effect of the sub-vertical joints may have an
increased effect particularly considering that the beam is under a combined flexural-compressive load. Such combined
effect potentially reduces the critical stress that causes pure compressive buckling. Therefore, it seems appropriate to
adopt an estimate between these two values, and further assessed by numerical analyses if the approximate analysis
results indicate buckling may be an issue. Compressive buckling issues will be further illustrated in the next case. An
additional mechanism that should also be considered is the effect of stress induced failure. It is generally accepted that,
for unconfined conditions, significant but stable micro-cracking starts at approximately 40% of the UCS of the intact
rock with unstable cracking typically initiating at 70% to 80% of the UCS. For instance, Figure 10c demonstrates that
for a thickness of 0.6 m the maximum compressive stress developed through arching is approaching the rock UCS.
Such cracking may deteriorate the rock modulus causing further beam deflections which could promote a transition of
the beam behaviour from a meta-stable/yielding condition to a collapse.
STABLE
140
120 UDEC
(a) Deflection (mm) Analytical
COLLAPSE 0
1 2 3
Shear
strength
reduction
factor
(c) φ = 7.8º, kn/ks = 1 GPa/m (d)
Figure 12: Deflection contours for weak abutment contact (a) FoS = 2.31 MPa (b) FoS = 1.5 MPa (c) FoS = 1 (d)
Impact of shear strength reduction.
The UDEC model presented in Figure 12a indicates a similar factor of safety to that of the analytical solution, i.e. a
stable condition. Figure 12d illustrates the development of the voussoir beam maximum deflection upon a shear
strength reduction up to a value of FoS=2.31 where collapse occurs (Figure 12c) at the abutment. Note that the
maximum beam deflection in Figure 12a is approximately only 10% higher than that depicted in Figure 6a indicating
that for FoS ≥ 2, the effect of a single weak abutment is less pronounced and beam deflections assessed with the
analytical solution are acceptable. Such results contradict the analysis carried out by Peck et al. (2013) and confirm the
ability of the analytical solution to provide a meaningful assessment for the case of a single weaker abutment, with a
difference in FoS of less than 5%. The effect of the horizontal stress case was not assessed as it was evident that it
would yield similar results to that of Figure 6c.
STABLE
50
UDEC
(a) 40
Collapse
0 10 20 30 40 50 Elastic
beam
30
STABLE 20
10
(b)
0
META-STABLE / YIELDING 0 5 10 15 20
Horizontal
stress
(MPa)
(c) (d)
Figure 13: Deflection contours weak zone within intact rock (a) σh = 0 MPa; (b) σh = 8 MPa; (c) σh = 16MPa; (d) Effect
of initial horizontal stress.
3m
Figure 14: Geometry of composite rock beam with seams and rock bolting
Figure 15 presents the beam deflections of two comparison models: with and without weak seams. The weak seams
increase the beam deflections by approximately 50% for all levels of stress but both are stable. It is important to note
that Figure 15c also demonstrates that, in this case, the rock bolts are not fully effective in reducing the overall beam
deflection with the composite beam behaviour very similar to that of a single 1 m thick bed. This is mainly due to both
rock joints and rock bolts being modelled at an angle of 90º to horizontal which results in small mobilisation of axial
and shear forces along the bolts for the current level of deflections. The impact and performance of rock bolts will be
further investigated in the next section.
STABLE
40
UDEC
-‐
3
x
1m
bed
with
seams
UDEC
-‐
3
x
1m
bed
without
seams
STABLE
10
0
0 5 10
(b) Horizontal
stress
(MPa)
(c)
Figure 15: Deflection contours for rock bolted beds (a) σh = 0 MPa without seams (b) σh = 0 MPa with seams (c) Effect
of seams and initial horizontal stress.
The rock bolts were modelled in UDEC using reinforcement elements which offer both axial and pure shear resistance.
The axial and shear stiffness, Ka and Ks respectively, were estimated by Equation 4:
3/ 4
⎡ ⎤
⎢ ⎥
G g Eb ⎢ 2Eg ⎥ (4)
K a = πd1 K s = Eb I
⎛ d ⎞ ⎢ ⎛ d 2 ⎞ ⎥
2⎜⎜ 2 − 1⎟⎟ ⎢ 4 Eb I ⎜⎜ − 1⎟⎟ ⎥
⎝ d1 ⎠ ⎣ ⎝ d1 ⎠ ⎦
where d1 is the rock bolt diameter (d1 = 21.7 mm), Gg is the grout shear modulus (Gg = 12 GPa), Eb is the Young’s
modulus ot the rock bolt (Eb = 210 GPa), d2 is the rock bolt hole diameter (d2 = 54 mm), Eg is the grout Young’s
modulus (Eg = 30 GPa) and I is the rock bolt second moment of area. The rock bolts shear capacity, Ps,ult were estimated
by:
2
Ps,ult = 0.67d1 Fbyσ ci (5)
where Fby is the ultimate stress of the rock bolts (Fby = 920 MPa) and σci is the uniaxial compressive strength of the
intact rock (σci = 23 MPa). The axial capacity was assumed Pa,ult = 340 kN or limited by a bond strength of 1 MPa. In
addition, an ultimate axial strain of 15% was set to the rock bolts beyond which both axial and shear capacity drops to
zero.
STABLE
(a)
Figure 16: Shear displacements along bedding partings and forces for rock bolts at 90º, σh = 0 MPa and no surcharge.
The low mobilisation of bolt shear forces observed in Figure 16 is related to the small beam deflections and consequent
small shear displacements mobilised along the bedding partings. In order to further investigate the performance of the
rock bolts, a triangular surcharge ranging from zero to q = 195 kPa mid-span is applied at the top of the model used in
Figure 16. This triangular surcharge is equivalent to a 7.5 m overburden above roof beam which is considered
reasonable for a span of 15 m even for depths of more than 10 m due to stress arching effects. As expected, greater
beam deflections promotes additional shear displacements which further highlights the positive effect of the rock bolts
at low values of horizontal stress (Figure 17).
Figure 17 also illustrates the significant improved effect rock bolts at a shallower angle, in this case 70º (Figure 20).
The impact of installing rock bolts at an angle can be clearly illustrated (Figure 18) through the analytical solution for
rock bolt design proposed by Pells (2002). For bolts at 90º, the analytical solution predicts a shear force of
approximately 45 kN mobilised at a shear displacement of about 3 mm whereas 300 kN is mobilised with less than 2
mm for an angle of 70º, and clearly explaining the significantly lower deflection of Figure 17c. It is important to note
that the UDEC reinforcement element only predicts 50% of the analytical solution bolt shear force at 90º. However,
similar results can be achieved if the UDEC parameters, particularly the shear and axial stiffnesses, are calibrated with
the analytical solution or perhaps based on laboratory bolt shear tests.
STABLE
60
UDEC
-‐
3
x
1m
bed
without
bolts
50 UDEC
-‐
3
x
1m
bed
with
bolts
@
90º
Beam
deflection
(mm)
Deflection (mm)
(a) UDEC
-‐
3
x
1m
bed
with
bolts
@
70º
0 10 20 30 40 50 40
30
STABLE
20
10
(b)
0
STABLE
0 5 10
Horizontal
stress
(MPa)
(d)
(c)
Figure 17: Deflection contours for bedded beam with triangular surcharge q = 195 kPa (a) σh = 0 MPa without rock
bolts (b) σh = 0 MPa with rock bolts at 90º (c) σh = 0 MPa with rock bolts at 70º (d) Effect of initial horizontal stress.
(a) (b)
Figure 18: Analytical solution for rock bolts (a) at 90º angle to bedding (c) at 70º.
(a)
Deflection (mm)
0 10 20 30 40 50
Figure 19: Deflection contours for single steeply dipping joint (dip 60º) with rock bolts at 90º and σh = 0 MPa.
A second rock bolt pattern is investigated. The locations of the rock bolts are exactly the same as in Figure 19 but
installed at an angle 70º. Such bolt pattern not only improves the performance of the rock bolts as discussed above but
also increases the chances of more bolts crossing steeply dipping joints. In this example, the number of bolts crossing
the joints increases to two as depicted in Figure 20. Note that to achieve the same number with a 90º angle, the bolts
spacing would have to be less than 1 m and they would still not have the same shear performance. Spot bolting is not a
practical option as such steep feature may not always be visibly evident in the roof during excavation.
15 m
3m
Figure 20: Geometry of composite rock beam with single steeply dipping joint.
Figure 21 presents the results of the UDEC model with rock bolts installed at a 70º angle. As expected, there is
significant improvement in the rock beam behaviour which becomes stable with a maximum deflection of
approximately 8 mm. Such smaller deflection results in lower axial strains in the bolts now able to sustain the applied
loads.
STABLE
3.5
10
UDEC 3.0
Yielding
Collapse
Beam
thickness
(m)
8
Beam
deflection
(mm)
1.5
4
1.0
2 0.5
(b)
(c)
0 0.0
0 50 100 150 200
0 1 2 3 4 5 6 7 8 9 10
Horizontal
stress
(MPa) Midspan
deflection
(mm)
Figure 21: (a) Deflection contours for single steeply dipping joint (dip 60º) with rock bolts at 70º and σh = 0 MPa (b)
Effect of initial horizontal stress (c) Analytical voussoir beam.
It is important to note that the analytical solution does not account for multiple beds, thus, not directly applicable for
composite beams. However, analysis is often carried out ignoring the bedding partings, thus, with a beam thickness
assumed to be 3 m. In this case, the analytical solution predicts a mid-span deflection in the order of 2.5 mm (Figure
21b and Figure 21c). The major difference in this case, i.e. 2.5 mm compared with the 8 mm of the UDEC model, could
be attributed to the single steeply dipping joint. However, as discussed above, the roof span to beam thickness is 5, i.e.
less than 10, which again invalidates the direct applicability of the analytical solution. In fact, if a 3 m beam is modelled
in UDEC without the bedding partings, the model predicts collapse of the beam due to sliding at the abutments which
seems to indicate that for low roof span to thickness ratios, the Diederichs and Kaiser (1999) parabolic function used to
assess the stress arching mechanism is not applicable. Interestingly, in this case, if the joints at the abutment are made
elastic in shear (infinite cohesion) but allowing for tensile failure (zero tensile strength), the UDEC beam deflection
matches the analytical solution.
12 m 7m
(a) (b)
Figure 22: Example of flat-roofed tunnel in Ashfield Shale (residual soil overlying Class IV Shale, Class II Shale and
Class II Sandstone): (a) bolts at 90º; (b) bolts at 70º.
8 CONCLUSIONS
The design of semi flat-roofed tunnels in Sydney using a voussoir beam analogue has been proven successful over time,
despite the theory limitations. However, if the geological conditions are clearly understood and good engineering
judgement is applied to take these conditions into account, the voussoir beam analytical solution can be confidently
used in practice to yield similar results to that of more comprehensive discontinuum numerical models. In other words,
if the appropriate mechanisms are well understood and considered in the analytical solution, as they should also be
accounted for in the numerical models, the predicted behaviours are in most cases similar.
As detailed in analyses presented above, rock bolts are typically used to stitch together near horizontal beds of limited
and variable thickness, to ‘trick’ the rock mass into behaving as an appropriately thick linear arch, robust against the
uncertainties in joint directions that may be expected in strata such as the Hawkesbury Sandstone of Sydney. In
addition, the existence of relatively high horizontal stresses are the second key feature in determining a geological
environment amenable to flat or semi-flat roof designs.
9 ACKNOWLEDGMENTS
The authors of this paper would like to acknowledge the contribution of Patrick Wong who reviewed the paper
providing useful comments.
10 REFERENCES
Asche, H.R. and Lechner, M.K. (2003). Design for the Cross City Tunnel, Sydney. Proccedings of the RETC. Seattle.
pp. 122-139.
Bertuzzi, R and Pells, P J N (2002a). Geotechnical parameters of Sydney Sandstone and Shale. Australian
Geomechanics, 37(5).
Bertuzzi, R and Pells, P J N (2002b). Design of rock bolt and shotcrete support of tunnel roofs in Sydney sandstone.
Australian Geomechanics, 37(3).
Booker and Best (1990). Coffey internal communication.
Diederichs, M S and Kaiser P K. (1999): Stability Guidelines for Excavations in Laminated Ground - The Voussoir
Analogue Revisited, Int. J. Rock Mech. & Min. Sci.; 36, pp 97-118.
Evans W H. (1941). The strength of undermined strata. Trans. Inst. Min. Metall., 50:475-500.
Itasca (2006). Universal Distinct Element Code User’s Guide.
Pells, P J N, Best, R J, and Poulos, H G (1994). Design of roof support of the Sydney Opera House underground
parking station. Tunnelling and Underground Space Technology, Vol 9, No 2, pp201-207.
Pells (1993) Rock Mechanics and Engineering geology in the Design of Underground Works. EH Davis Memorial
Lecture, Australian Geomechanics.
Pells, P. J. N. (2002). Developments in the design of tunnels and caverns in the Triassic rocks of the Sydney region.
International Journal of Rock Mechanics and Mining Sciences 39: 569-587
Peck, W A, Sainsbury, D P and Lee, M.F (2013). The importance of geology and roof shape on the stability of shallow
caverns. Australian Geomechanics, 48(3).
Sofianos A I (1996). Analysis and Design of an Underground Hard Rock Voussoir Beam Roof. Int. J. Rock Mech. &
Min. Sci., 33 (2), pp 153-166.
Abstract
The voussoir beam analogue has provided a useful stability assessment tool for more than 55 years and has seen numerous
improvements and revisions over the years. In this paper, a simpli®ed and robust iterative algorithm is presented for this model.
This approach includes an improved assumption for internal compression arch geometry, simpli®ed displacement determination,
support pressure and surcharge analysis and a corrected stabilizing moment in the two dimensional case. A discrete element
simulation is used to verify these enhancements and to con®rm traditional assumptions inherent in the model. In the case of
beam snap-through failure, dominant in hard rock excavations of moderately large span, design criteria are traditionally based
on a stability limit which represents an upper bound for stable span estimates. A de¯ection threshold has been identi®ed and
veri®ed through ®eld evidence, which corresponds to the onset of non-linear deformation behaviour and therefore, of initial
instability. This threshold is proposed as a more reasonable stability limit for this failure mode in rockmasses and particularly
for data limited cases. Design charts, based on this linearity limit for unsupported stability of jointed rock beams, are presented
here summarizing critical span±thickness±modulus relationships. # 1999 Elsevier Science Ltd. All rights reserved.
0148-9062/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 8 - 9 0 6 2 ( 9 8 ) 0 0 1 8 0 - 6
98 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
was ®rst proposed by Evans [11] speci®cally to explain The primary modes of failure assumed in the model
the stability of a jointed or cracked beam. After gener- and veri®ed in laboratory tests by Sterling [15] are
ating a great deal of controversy when ®rst published, buckling or snap-through failure, lateral compressive
the voussoir beam analogue has been generally failure (crushing) at the midspan and abutments, abut-
accepted and has since been reworked and presented ment slip and diagonal fracturing (Fig. 3). Shear fail-
as a simpli®ed tool for stability analysis of excavations ure (Fig. 3c) is observed at low span-to-thickness
in civil construction and in mining [10, 12±14]. Fig. 2 ratios (thick beams), while crushing (Fig. 3b) and
illustrates two example cases where the voussoir beam snap-through failure (Fig. 3a) is observed at higher
analogue can be invoked to explain the inherent stab- span-to-thickness ratios (thin beams). An examination
ility of a laminated hangingwall (Fig. 2a) and cross of the model data presented by Ran et al. [16] shows
jointed back (Fig. 2b) in hard rock environments. that if the angle between the plane of the cross-cutting
Fig. 2. Voussoir beams encountered at (a) Winston Lake Mine, Ontario and (b) mine access, Sudbury.
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 99
Fig. 4. Elastic beam with (a) ®xed ends and (b) simple (pin) supports.
100 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
beam, is found across all plane sections within the encountered in newly developed underground spans at
beam (Fig. 4a). The solution (Eqs. (1) and (2)) for the low to moderate depth. This initial elastic phase fol-
maximum stress values, at the abutments, for com- lowed by progressive fracture and deformation of
pression (bottom of beam) or tension (top of beam), laminated hangingwalls has been observed and is
smax, as well as the maximum beam de¯ection, d, can described in detail by Milne [18].
be easily calculated using closed form beam It is highly unlikely that any thinly laminated roof
equations [7]: structure will remain in a completely continuous and
elastic state after excavation. The transition from con-
gS 2
smax 1 tinuous elastic beam to voussoir beam is normally
2T
assured and assumed in most cases. Failure is not
inevitable in this situation, however. When the tensile
gS 4 cracking, described above, or lamination-normal joints
d 2
32ET 2 crosscut the beam and render it incapable of sustaining
where E is the Young's modulus of the rock and g is tensile stresses, a compression arch develops within the
the speci®c weight. The maximum stress at the mid- beam, rising from the abutments to a high point at
span is one half of the maximum stress at the abut- midspan. For a half-span, this arch generates a
ments. Therefore, for such a beam with ®xed ends and moment between the reaction force at midspan and at
distributed loading, yield is assumed when the maxi- the abutments (Fig. 5) which acts to resist the moment
mum tensile stress, in the upper part of the beam at imposed by self-weight. Horizontal stress symmetry
the abutments, exceeds the tensile strength of the rock. within the beam is lost, making a closed form solution
Vertical tensile fractures form at the abutments and impossible without assuming an average thickness NT
the beam becomes simply supported (assuming no slip for this arch. This is the case for the solution intro-
at the abutments), as shown in Fig. 4(b), with a maxi- duced by Evans [11] and later by Beer and Meek [12].
mum tensile stress at the midspan given by In both cases N is assumed to be 0.5. Numerical exper-
imentation by Evans showed that this was an incorrect
2gS 2 assumption and that the equilibrium value was closer
smax 3
3T to 0.7. Nevertheless, Evans chose the value, N = 0.5,
This stress is now higher than the previous abutment to simplify the practical solution which he was present-
stress, and therefore higher than the rock tensile ing. Investigations by the authors of this paper have
strength. This leads to subsequent fracturing centred shown that N is variable. While N is closer to 0.75 for
about the midspan as shown by Stimpson and stable beams at equilibrium, N drops to below 0.5 as
Ahmed [17]. This process of progressive cracking at the critical (unstable) beam geometry is approached.
the abutments, followed by cracking at the midspan In this formulation, the average thickness, NT, of
and other parts of the beam can be responsible for a the compression arch within the beam is initially
¯urry of low-level seismic emissions (rock noise), often unknown, as is the ultimate moment arm, Z0, between
Fig. 8. Flow chart for the determination of stability and de¯ection of a voussoir beam.
Z decrease, converging to a single rational solution The factor of safety with respect to crushing (com-
pair (N, Z) at the absolute critical limit. pressive failure) at the abutments and at midspan is
Finally. taking Z and Z0 (as determined using the given by the ratio of uncon®ned compressive strength
equilibrium value of N), the de¯ection, d, at the mid- of the rock with respect to the maximum compressive
stress calculated in the model:
span is simply Z ÿ Z0. The complete procedure for the
determination of stability and equilibrium de¯ection is UCS
F:S:crush 18
summarized in Fig. 8. fmax
104 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
The factor of safety with respect to vertical sliding, of In order to obtain a reasonable yield threshold for
an unsupported beam under self-weight, along joints the buckling limit, the relationship between midspan
at the abutments is given by displacement, d, and thickness, T, was considered.
Fig. 10 illustrates summary results for several beam
fmax N stinesses. The de¯ection/thickness relationships are
F:S:slide tan f 19 log-linear for beam geometries with ample thickness.
ge S
As the thickness is reduced (or the span is increased)
the relationship becomes non-linear and eventually
A numerical buckling limit, B.L., is introduced here becomes unde®ned at ultimate failure. Ultimate failure
which is the percentage of values of N within the occurs at a buckling limit of 100%, corresponding to a
range of 0 to 1 for which a solution (i.e. a real value displacement equivalent to approximately 0.25 T.
of Z) cannot be obtained. Fig. 9 illustrates the The onset of non-linearity consistently occurs at a dis-
decrease in normalized equilibrium arch thickness, N, placement equivalent to approximately 0.1 T. This
and the increase in buckling limit, B.L., with increasing `yield' point also consistently corresponds to a buck-
span/thickness ratio. N consistently drops below 0.5 as ling limit, B.L., of 35%.
the critical span/thickness ratio is approached and The yield threshold determined in Fig. 10 more clo-
reaches a limit of 0.35 immediately prior to failure sely corresponds to the critical displacement limit of
(snap-through). Ultimate collapse and the critical span approximately 0.15 T obtained by Mottahed and
versus thickness relationships proposed by Evans [11] Ran [19] through numerical modelling. Beyond this
and by Beer and Meek [12] correspond to a buckling displacement the jointed beam model used in their ex-
limit of 100%. In other words, stability is impossible if perimentation began to exhibit unstable behaviour and
there is no arch thickness which yields an equilibrium failure became imminent. This limit was independent
solution. For a more conservative approach, however, of span and the modulus.
a threshold can be speci®ed for the buckling limit in The choice of buckling limit aects the calculated
order to obtain design dimensions for the beam. critical span or the critical thickness (limiting case for
Fig. 10. Example determination of beam `yield' limit (example span = 20 m).
beam stability) as illustrated by the example in Fig. 11. the discussion which follows. Typically thin hard rock
These limits are obtained by including the analysis out- beams (signi®cant compressive strength) under their
lined in Fig. 8 inside a simple iterative bisection algor- own self-weight will tend to snap-through before they
ithm to ®nd the critical value of a speci®ed input crush (Fig. 12). If an otherwise stable beam is sub-
parameter (in this case, thickness) which gives a value jected to surcharge loading, however, as is the case in
(span) straddling the interface between stability and many of the laboratory experiments described in the
failure. literature, then crushing failure will result (as may
The displacement limit of 0.1 T (B.L. = 35%) is abutment sliding or diagonal cracking). The in¯uence
used forthwith, as a yield limit, due to its convenience of surcharge loading on the stability of the beam as
and to account conservatively for uncertainties in the well as on the failure mode is demonstrated in Fig. 13.
approach. This limit is important for monitoring appli- The transition line (snap-thru/crushing) indicates
cations since it is independent of the span and rep- which failure mode controls the critical limits as
resents a universal rule of thumb for determining the shown.
signi®cance of measured displacements. Displacements
of less than 10% of the eective lamination thickness
(determined using a borehole camera or from mapping 3. Field evidence of snap-through limit
data) can be con®dently assessed as being within the
elastic limit of the voussoir beam, independent of rock 3.1. Mount Isa Mine
modulus or eective span.
Fig. 11 indicates critical limits for snap-through fail- A limiting displacement (for linear voussoir beha-
ure of the beam. Crushing failure is also considered in viour) of 0.1 to 0.15 times the mapped bedding thick-
106 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
Fig. 11. Eect of buckling limit on critical beam geometry for `snap-thru' failure of a horizontal beam (Erm = 10 GPa, g = 0.03 MN/m3).
Fig. 13. In¯uence of surcharge loading on beam stability and failure mode.
Fig. 14. Hangingwall response to mining compared with voussoir limits. Data from Milne [18].
108 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
3.2. Winston Lake Mine this study is used here to illustrate the voussoir beam
behaviour of the chert hangingwall. Borehole camera
A second case study from Winston Lake Mine in data [24] is also used here to estimate the eective
Ontario, Canada is used here to demonstrate the utility thickness of continuous partings parallel to the hang-
and validity of the voussoir analogue. The typical ingwall as the wall de¯ects.
stope geometry is similar to that in the previous The rockmass quality as indicated by the index
example except that the stopes are smaller (20 m high) Q [25] is approximately 7 to 28 which, using the
and less steeply dipping (458 to 608) and in this case, relationship [26]
shallower (550 m). Winston Lake Mine uses a modi®ed
Erm 25 log10 Q 20
version of the longitudinal retreat technique called
modi®ed AVOCA, in which the stope is blasted in full indicates a rockmass modulus in the range of 20 to 35
height slices perpendicular to the strike of the ore GPa. The data set, collected from numerous sources
body. Rock ®ll is introduced by end dumping some contains a large scatter and according to variability
distance back from the active face, creating an open limits proposed by Barton [27] this rockmass quality
stope of variable open strike length. The ®ll is com- could have a modulus as low as 10 GPa and as high
pressible rock®ll and cannot be considered tight ®ll as 50 GPa. The lower modulus could apply if the rock-
until signi®cant closure has occurred. As a stability mass is relaxed due to undercutting as was the case in
analogue, therefore, the two dimensional beam is valid some of the stopes [28].
in this case. The displacement performance of the hangingwall
As previously illustrated in Fig. 2(a), the hanging- can be predicted by the voussoir analogue. The results
wall is composed of blocky chert with surface parallel for this case are shown in Fig. 16. The equilibrium dis-
jointing along foliation planes and two orthogonal sets placement is plotted as a function of lamination thick-
of cross-joints. Cablebolt support is installed from a ness and rockmass modulus. The critical displacement
remote drift to supply full coverage at the hangingwall. limits for yield (10% of the thickness) and for collapse
Fig. 15 shows the layout of drifts, stopes, cablebolts (25% of the thickness) are overlain on the plot. The
and instrumentation for an experiment investigating joints at Winston Lake were not fully continuous and
cablebolt performance in hangingwalls and the impact the eective lamination thickness, therefore, evolved as
of stress change [21, 22]. Extensometer data [23] from the hangingwall de¯ected and as joints propagated
Fig. 15. Plan layout and typical cross sections through study area, Winston Lake Mine.
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 109
Fig. 16. Measured displacements and observed lamination thicknesses (data points), compared with voussoir predictions, Winston Lake Mine.
along the foliation planes or connected with other ing to the stability of the hangingwall by applying an
joints. The borehole camera surveys [24] can be used eective support pressure.
to crudely estimate the eective beam thickness at any
point in space and time. The measured displacements
are plotted against the estimated lamination thickness
in Fig. 16 for speci®c points along dierent extens- 4. Numerical veri®cation of snap thru mechanism
ometer-borehole camera clusters (Fig. 15).
The results show that the rockmass behaviour corre- A discrete element model [30] was employed to
lates well with the predicted displacements. The data further verify the analogue. Other authors have per-
with higher beam thicknesses are from earlier in the formed similar analyses with mixed results. A recent
paper by So®anos [31] describes a simulation using
mining. As more partings form, the eective thickness
®nite elements which appears to show that the voussoir
decreases and the points move to the left on Fig. 16. It
model presented by Brady and Brown [13], which is
is interesting that while the earlier points correlate
similar to the model described here, poorly predicts
with a rockmass with low modulus, the later data
the simulated displacements and beam stresses, primar-
seems consistent with a stier rockmass. This can be
ily due to an over-prediction of the eective thickness
linked to two possible mechanisms.
of the compression arch at the midpsan and at the
The ®rst is that as the voussoir beam de¯ects the abutments. The numerical simulations of So®anos [31]
compression in the arch increases. The rockmass mod- predict a normalized arch thickness of less than 0.3 at
ulus of a jointed rockmass is normally pressure equilibrium and less than 0.1 near failure. These are
dependent [29] and therefore, in this case displays a low compared to the values 0.75 for equilibrium and
strain-stiening behaviour. In addition, plain strand, 0.3 to 0.4 at failure, predicted by the iterative
cement grouted cablebolts tend to require up to approach in this paper. Numerical simulations by Ran
between 30 and 40 mm of slip to develop their peak et al. [16] indicate an arch thickness more consistent
frictional capacity [14]. If the eective embedment with the predictions in this paper.
length of the strand, equal to the lamination thickness, Clearly, something is amiss. The answer appears to
is less than the critical length required to break the be the numerical boundary conditions and discretiza-
steel, the cablebolt can provide this capacity over rela- tion used in the model. It was found during the course
tively large rockmass displacements. It can be seen of this research that a UDEC model [30] composed of
that late in the experiment, as mining progresses rigid blocks and elastic joints (no tension) and a model
farther past the instrumentation sites, the de¯ection composed of deformable blocks and elastic joints (no
beyond 40 mm follows a more stable trend than that tension) both seemed to exhibit similar behaviour to
predicted by the model. Cablebolts must be contribut- the model of So®anos [31] if the beam had ®xed sup-
110 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
Fig. 17. Exaggerated displacement pro®le (top) for a typical UDEC voussoir beam simulation and contours of internal horizontal compressive
stress (bottom).
ports or was bounded by rigid blocks as abutments. potential in cases where lamination thickness and mod-
This creates problems with discretization which do not ulus is uniform.
seem to permit the formation of a stress triangle The beam was ®rst allowed to de¯ect elastically
(0 < 1) at the abutments. Instead, reaction forces are while maintaining a non-zero tensile strength within
concentrated at the bottom corner of the abutments the joints. If this initial elastic de¯ection is not per-
leading to over-prediction of the moment arm and in- mitted, sliding will occur before arch stresses are gener-
accurate de¯ection predictions. ated. It is assumed that excavation is a gradual process
The model ultimately used for this paper utilized de- such that beam deformations occur before the abut-
formable elastic blocks (discrete blocks with internal ments are fully liberated. After elastic equilibrium was
®nite dierence zones), elastic joints (no tension) and achieved, the joint tensile strength was set to zero and
¯exible (elastic and internally discretized) but very sti the beam was allowed to continue deforming until
abutment blocks. This provided better displacement either equilibrium or failure occurred. The joints had
compatibility at the abutments. The results of this no cohesion but had a frictional strength. The initial
model correlate very well with the voussoir analogue elastic deformation stage allowed lateral stress and
as will be discussed presently. Simulations included frictional strength to accumulate in the joints, particu-
single beams only since the purpose of the model was larly at the abutments. Without this stage the beam
a comparison with the analogue model. Hatzor and would simply slip past the abutments under its own
Benary [32] have recently examined the in¯uence of weight.
multiple layers with interbed friction. The single beam Several con®gurations were tested, using dierent
prediction represents the worst case failure initiation rock stinesses, Ei, dierent joint normal stiness,
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 111
Fig. 18. Predicted variation of horizontal stress along arch axis (ABC) compared with UDEC results (top) and predicted and simulated stress dis-
tributions at abutment and at midspan (bottom).
JKN, and joint spacings, sj. The rockmass modulus, bolic variation introduced in this paper. Again the
Erm for comparison with voussoir predictions, was cal- match is acceptable for engineering application of the
culating using the relationship: method.
Models of unit thickness (T = 1) with several dier-
1 1 1 ent rockmass moduli were analyzed with increasing
21
Erm Ei JKNsj span, in increments of 5 m, up to snap-through failure
(blocks fall and beam disintegrates). Crushing failure
The de¯ection pro®le and the internal stress distri- was not investigated in this analysis. Fig. 19 illustrates
bution are shown in Fig. 17 for one example simu- the calculated equilibrium displacements in the UDEC
lation. The parabolic nature of the compression arch is models and the associated voussoir predictions. The
immediately apparent. Fig. 18 compares the lateral modeled de¯ections match the predicted relationships
stress distribution at the abutment and at the midspan well, with the exception of a slight over-prediction by
with the assumed triangular distribution and the pre- the voussoir model at higher stinesses. In addition
dicted voussoir arch thickness (0.65 T). The distri- the UDEC model failed in each case immediately
bution at midspan correlates very well, while the beyond the critical span predicted by the voussoir ana-
numerically simulated abutment stress distribution is logue.
more non-linear, resulting in a higher maximum stress The maximum stress occurred inevitably at the bot-
near the bottom edge of the beam. In addition, the lat- tom edge of the beam at the abutment. The midspan
eral stress variation along a parabolic path, ABC, in maximum stress correlated fairly well with predictions
the UDEC model is compared with the assumed para- (Fig. 20) while the abutment stress was consistently
112 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
Fig. 19. Midspan displacements (data points) from UDEC beam simulations compared with voussoir predictions (T = 1 m).
Fig. 20. Maximum compressive stress in UDEC beams compared with voussoir predictions (T = 1 m).
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 113
Fig. 21. Simpli®ed relationships for rockmass modulus as a function of rock quality and con®nement.
higher. If this is real and not an artifact of the model- and sample disturbance. The results have been further
ling, it is unlikely to devalue the predicted crushing modi®ed here to provide convenient functions for
limit for the voussoir analogue, since the area over modulus based on rock quality and con®nement
which the increased stresses occur in the UDEC model (Fig. 21). These functions can be used to estimate
is quite limited and any resulting initial crushing rockmass modulus for the voussoir beam. For most
would be highly localized. These numerical results, excavations where the voussoir analogue is valid, the
therefore, give us con®dence that the voussoir model lower bound moduli should be used.
can be used to predict the de¯ection and stability of Finally the compressive strength must be speci®ed.
rock beams which form in laminated ground. Crushing failure is local in nature and therefore relates
to the strength of the intact rock. However, in accord-
ance with recommendations by Martin [34] regarding
the strength of intact rock in situ, one third to one
5. Stability guidelines half of the laboratory UCS is used for this analysis.
The summary design limits based on yield (buckling
The results of this work can be summarized into limit = 35%) are presented in Fig. 22. If a beam thick-
normalized stability charts for use in design of under- ness is speci®ed, then a critical span is obtained for
ground openings. Parametric modelling has shown snap-through failure and for crushing failure. The de-
that the snap-through stability limits for critical span sign span is the lesser of the two values. Likewise, if a
and thickness can be related to the normalized rock- span is speci®ed, the critical thickness is the greater of
mass modulus, E *, which is equal to the modulus the two.
divided by the eective speci®c gravity, S.G.*(E * = E/ Actual spans are rarely fully two dimensional in
S.G.*). Similarly, the limits for crushing failure can be nature. In fact the two-dimensional beam represents a
related to the normalized compressive strength lower bound model for determining critical span. An
(UCS* = UCS/S.G.*). The eective speci®c gravity, upper bound is obtained from the model for a square
S.G.*, for a beam with a dip of a can be obtained span. Again, Brady and Brown [13] derive a relation-
from Eq. (22). ship for a square span by assuming that the span fails
as four triangular panels. The weight of each panel
S:G:* S:G: cos a 22
creates a moment about the abutment edge:
In order to simplify the estimation of rockmass mod-
ulus based on rockmass classi®cation, the test data for gTS 3
MW 23
®eld modulus presented by Barton et al. [27] and by 24
Bieniawski [33] was reexamined by Hutchinson and
Diederichs [14] to account for the eect of con®nement which is resisted by a reaction moment
114 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
Fig. 22. Stability guidelines for jointed rock beams (tunnel spans). Eective speci®c gravity, S.G.* = S.G. cos a.
fmax NTSZ Combining Eqs. (23) and (25) yields a modi®ed re-
MR 24
2 lationship for fmax:
This equation, however, is in error due to the assump-
tion of a constant moment arm, Z, through all sections ge S 2
of the triangular panel. The de¯ection at the hinge fmax 26
6N Z Z0
point is zero and therefore, the moment arm must
vary from Z = Z0 ÿ d, at the midspan, to Z0 at the
corners. Eq. (24) must therefore be revised to Eq. (26) is then substituted for Eq. (13) in the original
analysis. One ®nal adjustment is necessary, however.
fmax NTS Z Z0 For the square span, Eq. (16), for the shortening of
MR 25
4 the beam must be replaced by
Fig. 23. Stability guidelines for jointed rock plates (square spans). Eective speci®c gravity, S.G.* = S.G. cos a.
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 115
1 ÿ nL 2 N erated by the active pressure created by mechanical
DL fmax 27
E 9 3 rockbolts) resulted in a small increase in stability,
while cohesive resistance (no slip) generated by fully
where n is the Poisson's ratio for the rockmass. grouted bolts (or rock bridges) had a signi®cant sta-
The results for the square span are summarized in bilizing eect.
Fig. 23. . These charts are only valid if no low to mid angle
A few limitations with respect to these charts should jointing is present. In this case or in the case of
be noted: thinly laminated ground, support is necessary in
order to create a reinforced beam. Rockbolts or
. These charts can be used with con®dence in lami- (preferably [1]) grouted rebar should be approxi-
nated ground where the lamination thickness is mately equivalent in length to the desired beam
known and where the modulus can be estimated thickness [10]. Assuming a lower-bound rockmass
with some degree of reliability.
modulus, a reinforced thickness can be estimated
. It is imprudent to rely on such an assumptive
from Figs. 22 and 23 which will ensure adequate
method for very large spans (>100 m for steeply
stability. A factor of safety of 1.5 to 2 with respect
inclined spans and >60 m for horizontal spans)
(i.e. use half the span or twice the thickness) to is
where other in¯uences, not considered here, may
adequate for most situations [14].
govern stability. Consistent excavation quality is
. This method is not suitable for poor rockmasses
also assumed. Over-blasting, or uneven excavation
surface geometry will negatively impact roof stab- with a low RQD rating (<50) and more than three
ility. joints sets.
. The assumption in these charts is that the joints are . This technique is designed to predict the onset of
rough enough to provide frictional resistance under roof instability. Thus only the ®rst (lowermost) lami-
low to moderate con®nement (i.e. no slickensides or nation is considered and not a composite beam
low friction coating) and that the span to thickness structure. This is based on the assumption that, pro-
ratios are greater than 10. Sliding failure along joints vided the beam thickness is stability of the whole
at the abutments or within the beam is not con- roof is controlled by the stability of this ®rst beam.
sidered. . The in¯uence of the gravity load component parallel
. It is assumed that there is no frictional or cohesive to inclined laminations is ignored in order to achieve
resistance along the interfaces between the lami- a tractable solution. This leads to the apparently
nations. This represents a worst case assumption reasonable conclusion that inclined layers (dip = a)
since such resistance increases the stability of the are signi®cantly more stable than horizontal layers.
beams. Snyder [1] showed that limited friction (gen- The de¯ecting beam must still achieve the same
Fig. 24. Examples of parallel weight correction to critical span for inclined beams.
116 M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117
Fig. 25. Corrected equilibrium midspan de¯ections for stable steeply dipping beams (dip = 808), accounting for parallel weight component.
moment equilibrium (between the perpendicular 100% for near-vertical beam (dip = 808).
weight component and the compression arch reac- Undoubtedly, the in¯uence of the parallel weight com-
tion force) regardless of initial (gravitational) com- ponent is more complex than described here, leading
pression within the beam. The beam does, however, to assymetrical beam de¯ection and other compli-
settle under this imposed loading. The eect of this cations. A program of numerical experimentation is
additional compression can be considered implicitly warranted to investigate this in¯uence and provide
by imposing an initial shortening of the beam, DdC, more accurate stability predictions for inclined beams.
equivalent to the compression due to gravity (paral- Such a program is beyond the scope of this paper.
lel to the beam) given by Eq. (28). The beam must
®rst close this gap before additional compression
can be created during beam de¯ection. This gap is 6. Conclusions
used to calculate a new initial moment arm in
Eq. (29). This value is substituted for Z0 in Eq. (17). An improved iterative approach to a classic ana-
logue for stability assessment of laminated ground has
S 2 g sin a been presented with several improvements and correc-
DdC 28 tions including an improved assumption for lateral
E
stress distribution and arch compression, the appli-
s
cation of support pressure and surcharge loading, sim-
3S 8 2 pli®ed displacement determination and a robust
Z *0 Z 0 ÿ DdC 29
8 3S iteration scheme.
A linearity limit or yield limit has been identi®ed,
Examples of the in¯uence of this correction on critical corresponding to a midpsan displacement of approxi-
span are shown in Fig. 24. Given the bevy of simplify- mately 10% of the lamination thickness. This displace-
ing assumptions already inherent in the model, the ment limit appears to be independent of rockmass
impact of this correction is of limited practical signi®- modulus and is a useful guideline for performance
cance for rockmass moduli greater than one GPa and monitoring and observational design. Beyond this dis-
can be neglected in the interest of producing the sim- placement, stability cannot be assured. The yield limit
pli®ed and uni®ed charts in Figs. 22 and 23. On the and the methodology in general has been veri®ed using
other hand, according to this simpli®ed correction, ®eld evidence and numerical simulations.
non-critical equilibrium de¯ections are signi®cantly This limit is used to present stability charts for de-
aected as in the examples shown in Fig. 25, with sign. These design charts have been normalized with
increases, at one half of the critical span, exceeding respect to eective speci®c gravity which is a function
M.S. Diederichs, P.K. Kaiser / International Journal of Rock Mechanics and Mining Sciences 36 (1999) 97±117 117
of rock density, excavation dip angle and surcharge [14] Hutchinson DJ, Diederichs MS. Cablebolting in underground
mines. Canada: Bitech, 1996. 406 pp.
loading or support pressure. Two failure modes are
[15] Sterling RL. The ultimate load behaviour of laterally con-
presented for thin laminations. These are snap-through strained rock beams. The state of the art in rock mechanics.
and crushing. Other failure modes apply to thick Proc. of the 21st US Symposium on Rock Mechanics. 1980. p.
beams and are not considered. Critical span of the 533±42.
beam is determined by the critical failure mode (mode [16] Ran JQ, Passaris EKS, Mottahed P. Shear sliding failure of the
which gives the lowest critical span). jointed roof in laminated rock mass. Rock Mech. Rock Eng.
1994;27(4):235±51.
More work is needed to account for the boundary [17] Stimpson B, Ahmed M. Failure of a linear Voussoir arch: a lab-
parallel component of weight in steeply dipping vous- oratory and numerical study. Can. Geotech. J. 1992;29:188±94.
soir beams. While geometrical limits for stability are [18] Milne D. Underground design and deformation based on sur-
marginally aected by the neglect of this component, face geometry. Ph.D. Thesis, Mining Department, University of
British Columbia, Canada, 1996.
equilibrium displacement predictions may be signi®-
[19] Mottahed P, Ran J. Design of the jointed roof in strati®ed rock
cantly in error for steeply dipping beams. based on the voussoir beam mechanism. CIM Bull.
1995;88(994):56±62.
[20] Milne D, Pakalnis RC, Felderer M. Surface geometry assess-
Acknowledgements ment for open stope design. Rock Mechanics: Proc. Of the 2nd
North American Rock Mechanics Symposium, vol. 1.
Rotterdam: Balkema, 1996. p. 315±22.
This research has been funded by the Natural [21] Maloney S, Fearon R, Nose J, Kaiser PK. Investigations into
Science and Engineering Research Council (Canada). the eect of stress change on support capacity. Rock support in
A special thanks is due to Doug Milne (currently at mining and underground construction. Rotterdam: Balkema,
the University of Saskatchewan) for supplying raw 1992. p. 367±76.
[22] Kaiser PK, Diederichs MS, Yazici S. Cable bolt performance
extensometer data for the Mount Isa case example. during mining induced stress change: three case examples. Rock
Thanks are also due to Sean Maloney of the support in mining and underground construction. Rotterdam:
Geomechanics Research Centre and to Winston Lake Balkema, 1992. p. 337±84.
Mine and Noranda Technology Centre. [23] Maloney SM, Kaiser PK. Stress change and deformation moni-
toring for mine design: a case study. Field measurements in geo-
mechanics. Rotterdam: Balkema, 1991. 481±90.
[24] Maloney SM, Kaiser PK. Field investigation of hanging wall
References support by cable bolt pre-reinforcement at Winston Lake Mine.
Research Report. Geomechanics Research Centre, Laurentian
[1] Snyder VW. Analysis of beam building using fully grouted roof University, Canada, 1993. 140 pp.
bolts. Proc. of the Int. Symp. on Rock Bolting. Rotterdam: [25] Barton NR, Lien R, Lunde J. Engineering classi®cation of rock
Balkema, 1983. p. 187±94. masses for the design of tunnel support. Rock Mech.
[2] BeÂtournay MC. A design philosophy for surface crown pillars 1974;6(4):189±239.
in hard rock mines. CIM Bull. 1987;80(90):45±61. [26] Barton NR, Lùset F, Lien R, Lunde, J. Application of Q-sys-
[3] Miller F, Choquet P. Analysis of the failure mechanism of a tem in design decisions concerning dimensions and appropriate
layered roof in long hole stopes at Mines GaspeÂ. 90th CIM support for underground installations. In: Bergmand M, editor.
AGM. Edmonton, 1988. Paper No. 120. Subsurface space, vol. 2. New York: Pergamon, 1980. p. 553±
[4] Milne DM, Pakalnis RC, Lunder PJ. Approach to the quanti®- 61.
cation of hanging-wall behaviour. Trans. Inst. Min. Metall. [27] Barton N. Application of Q-system and index tests to estimate
1995;105:A69±A74. shear strength and deformability of rockmasses. Proc. Int.
[5] Villasceusa E. Excavation design for bench stoping at Mount Symp. on Engineering Geology and Underground Construction,
Isa Mine, Queensland, Australia. Trans. Inst. Min. Metall. vol. 2. Lisbon: IAEG, 1983. p. II51±70.
1996;106:A1±A10. [28] Kaiser PK, Maloney S. The role of stress change in under-
[6] Beer G, Meek JL, Cowling R. Prediction of behaviour of shale ground mining. Eurock '92. Rotterdam: Balkema, 1992. p. 396±
hangingwalls in deep underground excavations. 5th I.S.R.M. 401.
Symposium. Melbourne, 1981. p. D45±51. [29] Barton NR, Bandis S, Bakhtar K. Strength, deformation and
[7] Obert L, Duvall WI. Rock mechanics and the design of struc- conductivity coupling of rock joints. Int. J. Rock. Mech. Min.
tures in rock. John Wiley and Sons, 1966. 649 pp. Sci. Geomech. Abstr. 1985;22:121±40.
[8] Hoek E, Brown ET. Underground excavations in rock. London: [30] Itasca Consulting Group. UDEC: Universal Distinct Element
Inst. of Min. and Metall., 1980. 527 pp. Code, Version 3.00. 1996.
[9] Fayol M. Sur les movements de terain provoques par l'exploita- [31] So®anos AI. Analysis and design of an underground hard rock
tion des mines. Bull. Soc. Indust. Min. 1885;14:818. voussoir beam roof. Rock Mech. Min. Sci. Geomech. Abstr.
[10] Corlett AV. Rock bolting in the voussoir beam: the use of rock 1996;33(2):153±66.
bolts in ground support. CIM Bull. 1956;LIX:88±92. [32] Hatzor YH, Benary R. The stability of a laminated voussoir
[11] Evans WH. The strength of undermined strata. Trans. Inst. beam: back analysis of a historic collapse using DDA. Int. J.
Min. Metall. 1941;50:475±500. Rock. Mech. Min. Sci. 1998;35(2):165±82.
[12] Beer G, Meek JL. Design curves for roofs and hanging-walls in [33] Bieniawski ZT. Engineering rock mass classi®cations. New
bedded rock based on `voussoir' beam and plate solutions. York: Wiley, 1989.
Trans. Inst. Min. Metall. 1982;91:A18±A22. [34] Martin CD. 17th Canadian Geotechnical Colloquium: the eect
[13] Brady BHG, Brown ET. Rock mechanics for underground of cohesion loss and stress path on brittle rock strength. Can.
mining. Chapman and Hall, 1993. 571 pp. Geotech. J. 1997;34:698±725.
International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
Abstract
Details are given of the analytical methods used to design rockbolt and shotcrete support for tunnels and large span caverns under
relatively low cover in the near horizontally bedded Triassic sandstones of the Sydney region.
The paper provides a concise description of the engineering geology of the Sydney sandstones because it is fundamental to tunnel
support design that a valid geological model be the basis of any analytical design. Equations are provided which allow calculations
of the lengths, density and capacity of rockbolts for support in this geological environment. The paper also discusses the design
approach adopted for support of tunnels and caverns which are at sufficient depth to generate compressive and shear failure of the
rock mass, so-called True Rock Pressure.
r 2002 Elsevier Science Ltd. All rights reserved.
The conception of a design of a new structure can Mittagong a thin horizon (o20 m) of interbedded
involve as much a leap of the imagination and as Formation shales and sandstones
much a synthesis of experience and knowledge as any
Hawkesbury maximum thickness 290 m, predomi-
*Tel.: +61-2-9874-8855; fax: +61-2-9874-8900. Sandstone nantly near-horizontally bedded sand-
E-mail address: mailbox@psmsyd.com.au (P.J.N. Pells). stone but with some laminite beds
1365-1609/02/$ - see front matterr 2002 Elsevier Science Ltd. All rights reserved.
PII: S 1 3 6 5 - 1 6 0 9 ( 0 2 ) 0 0 0 5 8 - 8
570 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
Narrabeen maximum thickness 700 m, comprising 2.2. Engineering geology of the Hawkesbury Sandstone
Group sandstones with claystone horizons
When viewed in vertical section the Hawkesbury
As shown in Fig. 5 the Wianamatta Group shales and Sandstone may be divided into three facies, namely:
the Hawkesbury Sandstone are the dominant near
surface rocks, and because the shales form an eroded sheet facies
cap it is in the Hawkesbury Sandstone that most of the massive facies
structures discussed in this paper are located. mudstone facies o5% of formation
P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587 571
The sheet facies comprises sets of cross-bedded strata clay matrix 15–30%
bounded by planar near horizontal surfaces, see Fig. 6. secondary quartz 4–10%
The units range in thickness from fractions of a metre to siderite (iron carbonate) 2–4%
greater than 5 m but are typically of the order of a
The average composition of the matrix clay is 55–75%
metre. The near horizontal bedding planes give this
kaolinite, 20–30% illite and the balance mixed-layered
facies a sheet-like appearance when viewed from a
clays.
distance. The cross-beds typically dip to the northeast.
The term massive facies was coined to covey the gross 2.2.1. Strength
aspect when viewed from a distance and does not mean The sandstone comprises medium grained quartz
wholly structureless at closer inspection. Frequently, grains; quartz overgrowth of the grains frequently
sandstone bodies of this facies have a discordant provides an interlocking structure and the development
erosional lower surface and a planar concordant upper thus of crystal faces imparts a glistening effect in
surface. Mudstone (or shale) breccia commonly occurs fractured faces.
within troughs at or above the basal surface, but clasts, Substance strength properties (nominally 50 mm
and in particular mudchips and mudflakes, can occur diameter core) of fresh or slightly weathered sandstone
dispersed throughout. are typically:
Petrographic analyses indicate that typically the
Saturated unconfined compressive strength ðUCSÞ
Hawkesbury Sandstone has the following composition:
¼ 25245 MPa;
detrital quartz grains 50–75%
lithic fragments 2–4% Saturated Brazilian tensile strength ¼ 223 MPa:
572 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
seams is not clearly understood but they provide the RQD values in the fresh or slightly weathered rock
major weakness within the Hawkesbury Sandstone typically lie in the range of 75–100%.
sequence.
Cross bedding. Cross bedding (also termed current 2.2.4. Faulting
bedding) is an ubiquitous feature of the sheet facies. Faulting in the Hawkesbury Sandstone is relatively
Cross bedding planes are often marked by the deposi- rare [3] but there are at least three 20–40 m wide zones
tion of flakes of mica, graphite, and carbonaceous comprising high angle normal faulting, oriented NNE,
matter. The cross bedding usually does not represent running through the central business district. The zones
planes of weakness in fresh or slightly weathered are spaced at between 300 and 600 m. In addition, low
sandstone. However, in moderately to highly weathered angle thrust faults, substantially confined to bedding
sandstone the cross beds can form surfaces of relatively horizons, are also found.
low tensile and shear strength (f0 E302351). The above structures play little part in the rock
mechanics discussed in this paper but serve as a warning
that the Hawkesbury Sandstone is not always a benign
2.2.3. Jointing medium.
The dominant joint set strikes NNE with dips ranging
from 651 to 901 east or west, depending on location
across the city. A secondary, orthogonal, set comprises 2.2.5. Regional stress field
near vertical joints. Measurements of the natural stress field have been
The joints have substantial horizontal continuity made on many projects in the Sydney region using
(typically greater than 20 m). Their vertical continuity hydrofracture, rock slotter and strain cell overcoring
is variable. Many of the joints terminate on bedding techniques. The results of these measurements have been
horizons and may have vertical continuities of the order presented in several papers including Enever et al. [4],
of 5 m. However, about 30% of the joints transgress Enever [5] and McQueen [6]. Unfortunately, many of
several bedding horizons and have vertical continuities the stress measurement compilations combine results
of between 10 and 30 m. Spacings of the dominant NNE from different geological units and include results
joint set range between less than 0.3 m to about 5 m, obviously affected by topography.
with an average of about 1.5 m. A pattern frequently The writer considers that, within the Hawkesbury
observed within the Sydney area is that the joints may Sandstone to a depth of about 150 m, the following
occur in swarms. This means that anywhere between equations represent an appropriate expression of the
three and ten joints may occur over a distance of a few natural total stress field:
metres, with there then being a substantial gap before s1 ¼ sNS ¼ 1:5 þ 1:2sv to 2:0sv MPa; ð1Þ
encountering further joints of the same set.
The orthogonal joints (ESE) have similar continuities
to the main set but with spacings in the range 6–20 m. s2 ¼ sWE ¼ 0:5s1 to 0:7s1 MPa; ð2Þ
574 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
The data in Table 2 show that: 3.3. Classification using the Q-system
* the RMR system is very insensitive to the intact In accord with the approach set out in Section 3.2 for
strength, a parameter which is very important in the the RMR system, Table 3 gives the Q-system classifica-
engineering behaviour of the Sydney sandstone, tion [10,11] for the range of Sydney sandstone, again
* the best sandstone classifies at the bottom of assuming a north–south oriented tunnel at a depth of
Bieniawski’s Class 2, indicating a fair to good rock 20–50 m.
mass—which is very conservative given the specta- Unlike the RMR system, the Q-value classification
cular unsupported vertical cuttings and unsupported provides good discrimination of the range of mass
openings made in this rock, and properties of the sandstones.
* the system does not discriminate well between the
sandstone grades encountered in the Sydney Basin. 3.4. Limitations for support design
Table 2
Classification of typical range of Sydney sandstone using RMR
I II III IV V
1 Intact strength 2 2 2 1 0
2 RQD 18 17 13 5 3
3 Discontinuity spacing 15 15 12 10 8
4 Discontinuity conditiona 25 22 20 10 10
5 Groundwater 10 8 8 8 8
6 Adjustment for orientation 5 5 5 5 5
Total RMR 65 59 50 29 24
Class 2 3 3 4 4
Description Good Fair Fair Poor Poor
a
This is difficult to assess because of the need to define the relative priorities of continuity, roughness, joint infill and joint wall condition.
576 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
Table 3
Typical classification of range of Sydney Sandstones using NGI Q-system (o50 m depth)
I II III IV V
1 RQD 90 80 65 25 5
2 Jn 2 4 4 6 12
3 Jr 3 3 1.5 1 1
4 Ja 1 1.0 2.0 3.0 6.0
5 Jw 0.8 0.8 0.8 0.66 0.66
6 SRFa 2.5 2.5 5.0 5.0 5.0
bolting at 2.6 m spacing with bolt lengths of 4.5 m (no CSIR to work on tunnel design, the most striking
mesh, no shotcrete). The actual support was similar to personal discovery was that there appeared to be no
the original recommendations and heavier than the appropriate analytical design methods for rock tunnel
updated guidelines. The first collapse comprised about support, certainly not as understood in the fields of
300 tonne of crown, the second about 20 tonne. After the structural engineering, hydraulic engineering and even
collapses the support was increased to substantially soil mechanics. There was a lot of talk about the ‘art’ of
greater density and capacity than would be indicated by tunnel design, but on close examination much of this
the Q-system. seemed to be educated guesswork. This is probably why
It may be argued that the Q-value adopted from the those working in rock tunnel support design enthusias-
site investigation boreholes was wrong. The writer has tically, and somewhat uncritically, adopted the RMR
reviewed this matter using logging of the actual and Q classification systems when they appeared in 1973
excavation. This suggests a Q-value of about 13 (i.e. and 1974.
half the original value). However, the support recom- The writer accepts that tunnel design is different from
mendations would be the same. many other engineering design processes. However, it
In early 2000 a major system of TBM driven tunnels can be performed on a scientific basis using an intimate
was completed on the north side of Sydney Harbour to blend of engineering geology, precedent, structural
store peak sewage flows. The scheme included about analysis and the observational method during construc-
6.5 km of 3.8 m diameter tunnel, 3.7 km of 6.0 m tion.
diameter tunnel and 3.5 km of 6.3 m diameter tunnel. Fortunately, the relatively simple geology of the
All the tunnels were in Hawkesbury Sandstone, with Hawkesbury Sandstone has facilitated the development
depths of cover ranging from about 20 m to about 80 m. of analytical design methods over the past two decades,
The initial primary support, comprising rockbolts and methods which have been demonstrated to work well in
mesh, was designed using the Q-system. The actual rail, road and water tunnels, and major caverns. Before
density of rock bolting (bolt per metre) which proved proceeding to discuss the analytical methods currently in
necessary to install, following inadequate performance use it is critically important to reiterate Lauffer’s [13]
of the initial design, ranged between 5 and 9 times the categorisation of tunnel support loadings, namely:
initial design densities.
(i) Loosening pressure: forces or pressures caused by
Based on the information given above it is the practice
the weight of blocks or prisms of loosened or
of the writer, and co-workers, not to rely on the
potentially loosened rock in the crown and side-
recommendations of the Q and RMR systems for
walls of a tunnel.
support design in the Hawkesbury Sandstone. The
(ii) Swelling pressure: caused by volumetric increases
approaches which are used are set out in the remainder
of clays, claystones or other rocks due to exposure
of this paper.
to the atmosphere under altered stress conditions.
(iii) True rock pressure: arising when compressive
stresses generated in the rock around the tunnel
4. Adopted support design methodology
are sufficient to cause compressive yielding and
fracture.
4.1. Basis of design
Swelling pressures are not an issue in Sydney’s
Some 30 years ago when, as a young civil engineer, the Triassic rocks. Loosening pressures are the prime
writer joined Bieniawski’s rock mechanics group at the consideration at depths of less than about 50 m, and
P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587 577
are discussed in Section 4.2. True rock pressure creates a [22]. This requires a decision to be made on what is the
much more difficult design problem and is discussed in allowable crown sag, and cognisance must be taken of:
Section 4.3. * the mass modulus of the crown strata and the
effective abutment stiffness, and
4.2. Design for loosening pressures * the natural stress field.
Fig. 7. Geometric disadvantages arising from cutting arch crown in Hawkesbury Sandstone.
578 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
Fig. 8. Theoretical design curves from linear arch analysis with examples from Sydney projects.
used to estimate these shear displacements, as used for local support of isolated loosened blocks of
discussed by Bertuzzi and Pells [24]. rock.
2. Calculate the shear stresses which would occur at the Fig. 10 shows the general case of a single rockbolt
locations of physical bedding horizons if behaviour crossing a discontinuity. The reinforcement acts to
were purely elastic. This can be done using the same increase the shear resistance of the joint by the following
finite element model but with elastic bedding plane mechanisms:
behaviour. The results for the simple example are
1. an increase in shear resistance due to the lateral
given in Fig. 9d. Again, if only horizontal bedding
resistance developed by the rockbolt via ‘‘dowel
discontinuities are considered, a closed form solution
action’’—force R1 ;
can be used for these shear stresses [24].
2. an increase in normal stress as a result of prestressing
Once the process of calculating the bedding plane of the rockbolt—force R2 ;
shear displacements and shear stresses is completed as 3. an increase in normal stress as a result of axial force
per (i) and (ii) above, attention can be turned to developed in the rockbolt from dilatancy of the
calculating the rock bolt capacities, orientations and joint—force R3 ; and
distributions required to create the effective linear arch. 4. an increase in normal stress as a result of axial force
This process is set out in Section 4.2.2. developed in the rockbolt from lateral extension—
force R4 :
4.2.2. Calculating rock bolting requirements The first component can be considered as increasing
At the outset it should be noted that consideration is the cohesion of the joint, while the other three
given here only to fully grouted rockbolts. These are components increase the frictional component of inter-
typically so far superior to end anchored bolts in their face strength by increasing the effective normal stress on
influence on rock mass behaviour that the latter are only the interface. If the rockbolts are at a spacing S; so that
P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587 579
where st is the tensile strength of the rock, sc the They appear to give similar predicted load versus shear
compressive strength of the rock, E the modulus of displacement curves to those obtained using the
the rock, f the friction angle of the crushed rock, Ty the equations presented in this paper. The difficulty is that
yield strength of the rockbolt, n the Poisson’s ratio of they include a ‘‘correction constant’’ which appears to
the rock, R2 the initial tension in the bolt, P0 the initial have no objective method of quantification.
stress in the rock in the plane of the joint (assumed equal
all round the bolt), d the shear displacement on the joint. 4.2.2.3. Rock bolt length. The bolt length is usually
Eqs. (7)–(10) can be solved, using a MathCAD routine taken as the required linear arch thickness plus 1 m. This
or a spreadsheet, to give a relationship between joint presumes there to be a physical bedding plane at the
shear displacement and the dowel action resistance R1 : upper surface of the nominated linear arch and is
intended to allow sufficient bond length for mobilization
4.2.2.2. Calculation of axial forces due to dilation and of bolt capacity at this postulated plane.
lateral rockbolt extension: forces R3 and R4. Calculation
of forces R3 and R4 is based on observations in 4.2.2.4. Bolt density. The design process is iterative
experimental tests by Dight [18], Pells [20] and Pellet because of the following variables in regard to the bolts
and Egger [21] that, for fully grouted rockbolts, axial alone:
bolt yield is attained at a joint after very small shear or * bolt capacity—a function of diameter and bolting
joint opening movements (typically o1.5 mm). material (typically either 400 MPa reinforcing steel or
The distance to the plastic hinges from the joint nominally 950 MPa steel associated with Macalloy/
surface is given approximately by the equation VSL/Diwidag bars),
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
D 1:7psy R2 2
* bolt inclination,
L¼ 1 : ð11Þ * bolt spacing across and along the tunnel.
4 Pu Ty
Typically, for tunnels of spans up to about 12 m use is
If the assumption is made that, under small shearing made of standard rockbolt steel (nominally 400 MPa).
displacements, axial strain in the rockbolt is dominantly For larger spans some, or all, of the bolts comprise high-
between the two plastic hinges, then from the geometry grade steel.
given in Fig. 10 the equations for R3 and R4 are, as a It is advantageous to incline bolts across the bedding
function of shear displacement: planes provided one is certain as to the direction of
d tan i pD2 shearing. However, as shown in the example presented
R3 ¼ Es ; ð12Þ
2L 4 in Fig. 9 such shearing is not always symmetrical about
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1 the tunnel centreline. Bolts inclined across bedding
0
L L2 ðd=2Þ2 pD2 against the direction of shearing can be largely
R4 ¼ Es @ A ; ð13Þ ineffective. Therefore given the uncertainty in this
L 4 regard it is considered appropriate that only those bolts
located over the tunnel abutments should be inclined,
where Es is the modulus of the rockbolt material (steel), the central bolts are installed vertically. As an example,
i the joint dilatancy angle. Fig. 12 shows the support used for the wide span section
If the bolt is inclined at an angle (p) to the joint such of the Eastern Distributor.
that forces in the bolt resist shear movement, then the Having made the above decisions regarding bolt
increased shear resistance due to R3 and R4 is lengths and inclinations the process of bolt density
Ds ¼ ðR3 þ R4 Þ sinðp þ d=LÞ tanðfj þ iÞ þ ðR3 þ R4 Þ computation proceeds, in principle, as set out below.
cosðp þ d=LÞ; ð14Þ
Step 1 The tunnel crown is divided into patches at each
where p is the angle of bolt to shear surface (radians). bedding horizon, as illustrated in Fig. 9, with
Eqs. (7)–(10) and (14) can be used to calculate the each patch intended to cover one rock bolt. It
total shear resistance provided by any rockbolt at any should be noted that the first major bedding
angle to a joint. The equations give good agreement with horizon above the crown usually controls design.
laboratory tests performed by Dight [18] and Pellet and Step 2 From the jointed finite element analysis (see
Egger [21], and with calculations made by Poulos [22] Fig 9b) the average shear displacement and the
using the analogy of a laterally loaded pile. Fig. 11 normal stress (see Fig. 9c) within each patch
shows the calculated combined forces R1 ; R2 ; R3 and R4 are calculated.
using the equations presented above, and similar Step 3 A rockbolt type (diameter, material, inclina-
predictions made by Poulos. tion) is selected for a patch and the forces
Equations to calculate the effect of fully anchored R1 2R4 are calculated as per the equations
rockbolts are also provided by Pellet and Egger [21]. given in Section 4.2.2.
P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587 581
Fig. 11. Calculated resistance forces R1 2R4 for typical full column grouted rockbolts.
these blocks will be located. In reality they will occur at Steel fibres Dramix 30 or 40 mm at about 50–
only a few locations in the crown of the tunnel, but 60 kg/m3
because the shotcrete must be applied in a pre-planned, Aggregate grading EFNARC [25]
systematic manner, and because safety requirements Microsilica about 5% of cement content
dictate that not even a brick size piece of rock may be
unsupported, it is necessary to assume that the ‘design Shotcrete designed for adhesion
block’ can occur anywhere. It comprises a patch of
gravity load on the shotcrete. UCS 35–45 MPa
Steel fibres Dramix 30 mm at 20 kg/m3
4.2.3.2. Shotcrete action. The shotcrete may be de- Aggregate grading EFNARC [25] or finer
signed to act in one of the two ways: Microsilica About 8–10% of cement content
(i) as a membrane spanning between bolts, or
4.2.4. Role of the observational method
(ii) by adhesion to the rock immediately around the
The design process described above cannot be done
design block.
for each metre of tunnel, and the practice is to develop
The two mechanisms are illustrated in Fig. 13. designs for the Typical, Adverse and Special conditions
Under membrane action the flexural strength of the expected in each Region of the geotechnical model
shotcrete is the key material parameter and steel fibres, prepared for the project, as described by Pells and Best
or mesh, are essential. In addition considerable attention [26]. Mapping and monitoring is essential during
has to be given to detailing a good structural connection construction to check that geotechnical conditions are
between shotcrete and rockbolt heads. Spider plates, as as expected and that crown deflections are consistent
illustrated in Fig. 14, were used for this purpose on the with design expectations.
Eastern Distributor project. The Opera House Carpark cavern crown, with its 6 m
The adhesion mechanism involves the shotcrete acting of rock cover, was designed using the principles
in shear, [23], and therefore flexural strength (and high described above and crown sag was carefully monitored
dosage of steel fibres) is of little relevance. The key as the 17.5 m span was progressively created by stripping
factor for ‘adhesion’ behaviour is that the rock surface is from an initial 6 m heading. Fig. 15 shows how crown
very clean. centreline sag increased as the span was progressively
Current practice in the Hawkesbury Sandstone is to increased, compared with theoretical predictions made
use the adhesion mechanism when designing in Class I at design stage.
or Class II sandstone, and to use the membrane concept Maximum crown sag measurements from four differ-
in poorer quality sandstone and shales. ent projects are shown in Fig. 8 compared with
Details of the methods of calculation are given by theoretical ‘‘linear arch’’ predictions.
Bertuzzi and Pells [24] and Barrett and McCreath [23] An important observation which has been made from
and are not repeated here. extensometers installed in the crowns of the Eastern
Distributor and M5 tunnels is that it is impractical to
4.2.3.3. Typical shotcrete specifications. Current design install rockbolts sufficiently close to the advancing faces
specifications in the Sydney tunnels are typically as of initial headings to prevent significant opening of
summarised below. bedding planes which exist less than about 0.5 m above
Shotcrete designed for membrane action the crown. With opening of such bedding planes close to
the crown it is not possible to generate the shear
UCS 40–50 MPa response equivalent to a pseudo-clastic beam, as
Residual flexural 2 MPa at 2 mm required by the design method given in Section 4.2.2.
strength (ASTM) To allow for this reality it is recommended that the
Fig. 14. Spider rockbolt plate used on Eastern Distributor to ensure adequate connection between shotcrete and rockbolts.
Fig. 15. Prediction & measurement of roof sag of the Opera House Carpark cavern.
lower 0.5 m of crown be treated as dead load (15 kN/m) manifests as loading imposed on crown support by
and not be included as part of the design linear arch compressive, shear and tensile failure of the near
thickness. horizontally bedded rock.
Design for true rock pressure is more difficult than for
4.3. Design for true rock pressure loosening pressure because a coherent analytical design
method is yet to be developed. However, the principles
Given the relatively high horizontal stress in the of design are quite clear, and safe designs for deep
Sydney Basin (see Section 2.2.5), true rock pressure tunnels and caverns in the Sydney Basin rocks
584 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
can be developed as outlined below. The important steps periphery exceed the mass strength of the sandstone
are: (typically about 20 MPa as discussed in Section 2.2.1).
However, good evidence has been presented by Stacey
(i) Assess the likely locations and extents of rock mass
[27] and Martin et al. [28] that the onset of, and the
yielding.
extent of, brittle rock mass failure can better be estimated
(ii) Take steps to reduce the extent of the problem by
using a tensile strain criterion. The writer’s current
altering the position and shape of the excavation.
practice is to use a stress versus strength criterion for
(iii) Recognise that no reasonable amount of support
assessment of failure likelihood, and for selecting an
can prevent the rock mass yielding and that
appropriate excavation shape. The tensile strain criterion
support is there to contain the failed material and
is used to assess the likely volume of failure once it is
maintain the geometric integrity.
known that true rock pressure cannot be avoided. In
(iv) Make use of the observational method by appro-
Class I and Class 2 Hawkesbury Sandstone it is reason-
priate instrumentation, and implement the princi-
able to adopt a critical tensile strain of about 0.0005.
ples of NATM.
As discussed by Pells [29], crown stress concentrations
Steps (i) and (iii) are dealt with in more detail in the under a high horizontal stress field are substantially
following sub-sections. Step (iv) is dealt with in many increased by the presence of low shear strength and/or
current texts and there are no special features applicable low stiffness near-horizontal bedding discontinuities.
to the Sydney Basin rocks. This is illustrated in Fig. 16 for the cases of a circular
TBM tunnel.
4.3.1. Location and extent of crown yielding If, at excavation level, the natural stresses are
Usual practice is to assume that crown yielding
initiates when the peak induced stresses at the tunnel sh ¼ s1 ¼ K0 sv ¼ K0 gd;
Fig. 16. Effect on low stiffness bedding horizons on stress concentrations around TBM tunnel.
P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587 585
where sh is the horizontal stress, s1 the major principal Clearly, particular site conditions affecting rock
stress, sv the overburden pressure, K0 the constant strength, bedding plane defects and topographic effects
which is calculated from Eqs. (1)–(3), d the overburden on the stress field must be taken into account in a
depth, g the unit weight. detailed assessment, and it should be noted that lengths
Then, in general, the maximum compressive crown of tunnel below 100 m in massive sandstone have
stresses can be expressed as shown no signs of failure. For flat-crown tunnels the
stress concentrations may be less than for a circular
s0 ¼ ðSf K0 1Þgd; tunnel [29].
where Sf is a constant depending on the shape of the
4.3.2. Bolting and structure requirements
tunnel and the anisotropy of the rock.
The characteristics of rockbolts appropriate for true
For a circular tunnel in isotropic rock, the value of
rock pressure are in some ways quite different from
Sf ¼ 3: Therefore, for g ¼ 0:024 MN/m3 the induced
those required for loosening pressures. As already
stresses (s0 ) would exceed a rock strength of 20 MPa at
stated, there is no way bolting can prevent yielding of
an overburden depth of 166 m. Actual experience in the
the rock. Therefore bolts must be able to accommodate
Hawkesbury Sandstone is that crown and invert failure
potentially large shear movements along joints and new
has developed around circular tunnels at depths of as
fractures, while at the same time providing high shear
little as 60 m. Invariably such failures are observed to be
resistance. Again full column grouted bolts are far
associated with low shear strength bedding surfaces just
superior to end anchored bolts.
above crown or just below invert (see Fig. 17). There is
The following types of bolts are ineffective or, at best,
no doubt that failure at such relatively shallow depth is
inefficient:
due to the stress considerations being similar to those
indicated by Fig. 16. * glass fibre reinforced plastic—because the long-term
For rapid ‘‘back-of-the-envelope’’ assessments the strength of this material is only about 30% of the
writer uses a value of Sf ¼ 6 for TBM tunnels, where short-term strength, and high local shear strains can
low-strength bedding planes are likely to be present, easily result in the long-term strength being exceeded
indicating that significant true rock pressure problems without the designer having any control of this
may have to be dealt with below a depth of about 75 m. matter.
Fig. 17. Typical stress induced failure in TBM tunnel of North Side Storage Project.
586 P.J.N. Pells / International Journal of Rock Mechanics & Mining Sciences 39 (2002) 569–587
* thin wall hollow steel bolts—because they provide standing of the important facets which must be covered
much less shear resistance than solid bolts. by a design, and the paper provides guidelines in this
* high tensile strength steel bolts—because strain to regard for rock masses similar to the Hawkesbury
failure may be as low as 5%. Sandstone.
[22] Pells PJN, Best RJ, Poulos HG. Design of crown support of the [26] Pells PJN, Best RJ. Aspects of primary support design for tunnels
Sydney Opera House underground parking station. Tunnelling in the Sydney area. Trans Inst Eng Aust 1991;CE33(2):57–66.
Underground Space Technol 1994;9(2):201–7. [27] Stacey TR. A simple extensions strain criterion for fracture of
[23] Barrett SVL, McCreath DR. Shotcrete support design in blocky brittle rock. Int J Rock Mech Min Sci 1981;18:469–74.
ground: towards a deterministic approach. Tunnelling Under- [28] Martin CD, Kaiser PK, McCreath DR. Hoek–Brown parameters
ground Space Technol 1995;10(1):79. for predicting the depth of brittle failure around tunnels. Can
[24] Bertuzzi R, Pells PJN. Design of rock bolts in tunnels in Sydney Geotech J 1999;36:136–51.
Sandstone. Sydney: ITA World Tunnel Congress, 2002. [29] Pells PJN. Geometric design of underground openings for high
[25] EFNARC. The European specification for sprayed concrete, horizontal stress fields. Proceedings of the Third Australian–NZ
Hampshire, UK, 1996. Geomechanics Conference, Wellington, NZ, 1980. p. 2–183, 2–188.