Você está na página 1de 8

A guide to safe and cost effective spillways

Dr Peter J Mason
MWH Ltd
Terriers House
201 Amersham Road
High Wycombe
Bucks HP13 5AJ
UK

Introduction
Hydropower dams will all require means of dissipating the energy of any surplus flows. As a minimum these
may represent a power station by-pass facility for use in the event of a sudden load rejection. The spillway and
energy dissipation capacity installed to achieve that will correspond to the installed station capacity with the
same amounts of power and energy to be dissipated. There may also be a need to pass amenity or surplus flows
when one or more turbines are inoperative.
More significant energy dissipation requirements will generally be required to safely discharge surplus flood
water. This will relate not to turbine capacity and flows but rather to those of major flood events. For example
in the case of Mangla dam in Pakistan the installed capacity of the power station is 1,000 MW whereas at times
of peak floods the power dissipation requirement of the main flood spillway is approximately 25,000 MW.
There are various standardised types of stilling basins and other energy dissipaters and these are listed in various
textbooks and design guides. However the guidance is generally centred around hydraulic performance based on
model testing. Their practical large-scale application, taking into account operational problems which may have
occurred with them over the years, is often not covered with the same thoroughness. This paper attempts to go
beyond theoretical hydraulic performance and guide the reader in terms of safe, while at the same time
economic, practical application. For brevity the paper focuses on the requirements of energy dissipating
structures rather than those of upstream conveyance works such as weirs, intakes and chutes. An exception is
made where chutes are designed to act as energy dissipaters.
Energy dissipation for large flows typically relies on one of the following general methods: Stepped or baffled chutes
Rock basins
Simple stilling basins including cascade basins
Baffle basins
Roller buckets
Flip buckets and plunge pools
Crest splitters and impact basins

Stepped and Baffled Chutes


Both stepped and baffled chutes dissipate energy as the flows pass along them. Thus, when the flow eventually
reaches tailwater the residual energy left to be dissipated is minimal. Stepped chutes follow two distinct flow
regimes. The first is referred to as nappe flow and the second as skimming flow. In the first the water passes
from step to step as a series of small waterfalls, dissipating energy as it does so. In the second, at higher unit
flows, the water "skims" over the steps at high velocity, trapping hydraulic rollers between each step. Much
more residual energy remains in this form of flow although some is still dissipated along the chute. More
importantly significant pressure differentials can build up between adjacent parts of the steps and the chute
walls. This can lead to severe damage and even collapse in the case of masonry spillways, see Fig 1.
Stepped chutes have become increasingly popular for use on roller compacted concrete (RCC) dams, where the
nature of the construction naturally suits their formation. They certainly have a reliable track record up to unit
flows of approximately 12 to 15 m3/s/m.

Figure 1 Major damage to the stepped masonry spillway at Boltby dam in the UK
Baffle chutes too dissipate energy over the whole chute and feature baffles covering the entire chute surface.
Well established guidelines exist for baffled chutes and successful large-scale applications include the
Driekloof dam spillway in South Africa and the 1,600 m3/s emergency bypass spillway at the Ghazi-Barotha
power station in Pakistan. They are effective but best used to fit specific needs as the limited unit flows, large
area and the large numbers of reinforced concrete baffles required can make them an expensive option. The
baffles may also be prone to damage where significant amounts of large floating debris are involved.

Hydraulic Jump Basins


The use of hydraulic jumps for energy dissipation includes rock basins, simple stilling basins, cascade basins
and baffle basins. Standardised guidelines, such as those used by the United States Bureau of Reclamation
(USBR) are available to proportion such basins, generally in terms of the Froude number of the incoming flow.
The basins will also require sufficient conjugate tailwater depths to force the jump to occur and to stop it
sweeping out of the basin. Both Froude number and conjugate depth requirements are fairly straightforward to
calculate. However operational experience has revealed at least three distinct types of problem can occur with
such basins: Under-slab pressurization and displacement
Abrasion due to ball mill action
Cavitation damage to baffle blocks.
A survey carried out by the writer some years ago, Mason (1982), indicated some generally safe operational
ranges for these types of basins and these are shown in figure 2. Broadly speaking, simple hydraulic jump
basins operate most successfully where the head difference between upstream reservoir level and downstream
tailwater level is less than 50m. Above that there have been a notable number of basin problems either in terms
of lifted slabs or abrasion due to debris circulating in the pool. This does not necessarily mean such basins
cannot successfully operate at higher heads, only that special care may be needed with their design, detailing
and operation.
Baffle basins seem to work best at upstream to downstream head differences of between 10 and 30 m. Within
that head range the baffles will intensified the jump, increasing the efficiency of energy dissipation and hence
reducing the required basin size. At below 10 m the Foude number of the incoming flow is insufficient to form
an effective jump and the energy is dissipated in swirls and eddies and at quite a low efficiency. Baffles are
ineffective at increasing the turbulence levels of such flows. For this reason, at head differences below 10 m,
simple basins are generally preferred. At upstream to downstream head differences of above 30 m the
associated velocities have in the past caused significant damage to protruding baffles and splitter teeth, see
figure 3, and so for upstream to downstream head differences of between 30 and 50 m, again simple basins are
preferable.

Figure 2 - Typical safe ranges of use for various forms of energy dissipater

Figure 3 Cavitation damage to splitter teeth at the high-head, Pit No 7 dam in the USA
Cascade basins are especially useful where a considerable length of gravity dam can be devoted to overspilling. Simple collector basins along the downstream toe of such an arrangement can collect the flows and
discharge them laterally in the forms of cascades, down the valley flanks and in towards a central main basin.
The increased hydraulic turbulence involved in such an arrangement can make it especially efficient
irrespective of the need for baffles.

Roller Buckets
Good guidelines exist for the design of roller buckets, both solid and slotted. Where deep tailwaters are
involved they can be a cost-effective option in place of conventional hydraulic jump stilling basins. However,
because the energy dissipation takes place immediately downstream and over foundation material, there have
been cases of damage due to circulating and entrained rock. This problem is accentuated if the buckets have
operated symmetrically for any reason. It is probably true to say that the use of such buckets has been less
popular in recent years.

Flip Buckets and Plunge Pools


For high head flows flip buckets and plunge pools are generally a preferred option as costs are minimised and
there is plenty of good and safe operational experience. Figure 2 indicates the typical head and flow ranges
where such prototype arrangements were found to operate successfully in the survey on which the figure was
based.
Conceptually the flow is conveyed to such buckets or ski jumps with minimal energy loss. The bucket then
deflects the flow as a free jet to some point downstream where it can excavate a scour hole or plunge pool in the
downstream river bed. The maximum projection length for such jets is achieved when the projection angle is
45 to the horizontal. However economies in bucket size can be achieved in practice by reducing the angle to
35 and with very little difference in projection distance. Jet impact angles of below 25 are generally not
preferred as the jet may fail to plunge fully and instead generate high lateral return currents.
The present writer has already given guidance on design aspects associated with such buckets and also with
associated downstream plunge pools, Mason (1993) and so these will not be repeated in detail here. Broadly
speaking the bucket radius should be approximately 5 times the flow depth in order to avoid streamline
separation. However some tightening of the exit radius to, say, 3 times the flow depth can be useful in shearing
the flow to increase jet turbulence. On some ski jumps, such as at L Aigle in France, the ski jump chutes are
super-elevated causing the flow to spiral and again leading to internal turbulence. Jets from counter opposing
ski jumps can also be made to collide prior to impact, increasing energy dissipation in the air. This may also
reduce scour depths. One disadvantage of such increased turbulence can be the generation of excessive spray.
This in turn can lead to adjacent hillside instability through saturation and/or affect adjacent electrical
installations such as switch yards.
In practice where the buckets and plunge pools have given rise to problems, studies have shown that these have
tended to be related to either: Asymmetrical gate operation in the case of wide chutes
Variable geology in plunge pool, again leading to asymmetrical scour development
Greater use than hitherto planned due, for example, to prolonged power station outages
Fortunately such problems are rare and very much the exception rather than the rule. One query that arises
quite regularly is whether or not to pre-excavate the plunge pool. Some prefer this on the basis that the pool
geometry is then defined and controlled. Certainly if the rock can be used elsewhere, such as for rockfill or
concrete aggregate, then pre-excavation may be useful, cost-effective and justified.
In the absence of that, where the rock would simply go to waste, it may be preferable to simply let spillway
flows excavate the plunge pool in due course. In such cases a rock bar will form downstream, raising the
tailwater levels. This may be useful in minimising further scour but will be detrimental where such tailwater
levels also affect power station generation. In such cases it is best to either site the plunge pool upstream of the
power station tailrace, or in a separate channel to it, or to make some on-going maintenance allowance to ensure
periodic rock removal to maintain generation capacity. Some degree of pre-splitting the foundation rock in the
plunge pool area is another option to guide the development of the subsequent scour.

Crest Splitters and Impact Basins


A very effective, but not widely known, method of high-capacity and high head energy dissipation is by the use
of crest splitter arrangements in conjunction with a downstream impact basin. Such arrangements were
originally developed in the 1930s and are often referred to as "Roberts splitters" in recognition of the man who
pioneered their use. They have been used successfully on many dams since, see figure 4.

Figure 4 Roberts type crest splitter teeth and deflector lip with downstream impact basin
at the Wadi Dayqah dam in Oman
They are particularly suited to concrete dams where the overspill weir is fitted with both downstream deflector
teeth and, below those, a continuous deflecting lip. In such a location, water velocities are not high enough to
cause cavitation concerns. The combination of teeth and lip breaks up the flow at high level such that the
falling flow is projected downstream in a less concentrated form than that of a solid jet.
Impact basins receiving the dissipated falling jet can either be unlined rock or, more generally for high heads,
concrete basins. In spite of the high heads sometimes involved there seem to be no reported problems with the
use of these basins, possibly because the flow is so broken and aerated prior to impact. Furthermore the depth
of tailwater required in the basin to avoid flows "sweeping out" is far less than would be the case with the same
flow entering a conventional hydraulic jump basin. Successful uses of this arrangement have included the
108m high Vanderkloof dam in South Africa, the 122 m high Victoria dam in Sri Lanka and the 75m high Wadi
Dayqah dam in Oman. Associated design flows for these were 8,500, 9,500 and 13,500 m3/s respectively.
Wadi Daydah is an RCC dam and Roberts splitters were used as an alternative to downstream steps in view of
the high unit flows involved.

Dissipater Selection and Efficiency


The appropriate form of stilling basin or hydraulic energy dissipater is often selected on the basis of personal
past experience and preference. However it is recommended that reference can, and should, also be made to the
guidance provided in figure 2. Another useful consideration is how efficient the basin might be in safely
dissipating the power and energy of the flow. Clearly a form of dissipation which has a high value of power in
relation to retained water involved will be more efficient, and so should require a more compact and less
expensive construction.
The approximate power densities of the various types of dissipater already listed are given below:

Simple hydraulic jump basins


Hydraulic jump basins with baffles
Simple cascade basins
Crest splitter teeth impact basins

5 to 10 kW/m3
15 to 20 kW/m3
30 to 40 kW/m3
50 to 60 kW/m3

By comparison plunge pools in rock below flip buckets have been shown to, typically, have a power density of
around 7 to 9 kW/m3 as they approach stability. However, the fact that construction costs associated with these
are limited to just that of the concrete flip-bucket, means that they are particularly cost-effective.

Other Factors Limiting Total Dissolved Gases


All the above discussions focus on the hydraulic efficiency and performance of various basins and other forms
of energy dissipater. In all cases aeration is involved as part of the dissipation process. Indeed, it can be argued
that the more efficient the dissipater the higher the degree of aeration. Increasingly engineers are required to
consider the environmental aspects of their work and under some circumstances such high levels of aeration can
be a disadvantage. In the case of large energy dissipaters, with stilling basin depths of 20 m or more, the
combination of depth and aeration can cause significant levels of gas to go into solution. Total Dissolved Gas
(TDG) saturation levels of 120% are generally accepted as an upper safe limit for downstream fish. Levels
much above that can result in large-scale fish kills as gases come out of solution inside the fish. The
phenomenon is the same as that experienced by divers who depressurised too quickly and develop the bends.
On the Rio Parana in Argentina, large fish kills occurred downstream of the Yacytera dam due to supersaturation of dissolved gases in a particularly large spillway flow. Gases went into solution due to the incoming
jet plunging down 20 m in the downstream stilling basin, but the process of gases coming out of solution was
much slower. Measurements of 150% saturation immediately downstream of Yacetera stilling basin still
corresponded to values of 107 to 110%, 90 km downstream. A significant factor was the wide and shallow
nature of the Rio Parana downstream of Yacyreta. The fish had nowhere to dive and recover. Where rivers
have much greater depths under flood the fish have places where they can stay deep and de-compress
gradually. Eliminating artificial aeration on the spillway chute was tested and had only a small effect (a
reduction of approx 7 to 10% in TDG levels). In the event, high level deflectors were tested and added to the
main branch spillway to stop the jet plunging. This was effective at low flows but high flows still plunged.
The design was a compromise between conflicting requirements, but generally successful. The downstream
profile of the piers was also changed to reduce the rooster tail potential for aerating the flow.
In practice it can be difficult to avoid this situation in the case of very large floods and shallow downstream
rivers. Indeed the phenomena often occurs naturally downstream of waterfalls. Current mitigation practice in
the US is to design spillways so that the one in 10 year flood, occurring over seven days, does not cause TDG
levels in excess of 120%. One way of achieving this is to limit stilling basin flow depths to not more than 8 to
10 m. This avoids the TDG levels associated with higher pressures. However in practice such basins will also
need to be sized for much larger flows. Limiting flow depths within them during low flows may not be
practicable. An alternative, which has some successful precedent, is to deflect the flows horizontally at
tailwater level to ensure that low flows enter only the upper part of the stilling basin water. Figure 5 shows the
model testing of such a deflector lip on the downstream slope of a conventional spillway. It can be seen that the
associated aeration in the stilling basin is mainly limited to the upper regions of the retained water. More
interestingly when the same basin was operated at full flow the effect of the lip seemed to improve the
dissipation as compared to that of a conventional basin.

Figure 5 Deflected flow limiting air entrainment to the upper zones of water
in a conventional stilling basin

Structural and Detailed Design


Obvious factors in terms of structural and detailed design include the hydraulic and hydro-dynamic pressures to
which various parts of the stilling basin or stilling device will be subject. Associated aspects include the
location and details of any joints, water bars, under-drains and anchor bars.
A typical hydraulic jump stilling basin fed from an upstream, sloping spillway chute will experience centrifugal
pressures at the upstream juncture between the chute and the basin. Flow depths may be small but the effective
water density will be high due to the change in flow direction. Pressures will then drop in that upstream part of
the basin subject to only to the depth of the incoming flow but rise to tailwater levels at, and immediately after,
the hydraulic jump. Aeration depths will also tend make effective water depths in the basin higher than
indicated in model tests. High aerated flow depths acting on the spillway sidewall will exert the same total
loading on the wall as an equivalent, non-aerated flow. However the higher aerated flow depths will exert a
greater bending moment on the wall and should be assumed as the design case.
Joints are best not located in the upstream zone subject to high centrifugal pressures. This avoids the possibility
of these pressures being transmitted elsewhere. On small, river barrage type structures it is not uncommon to
perforate stilling basin apron floors so that the pressures above and below the apron can locally balance.
However, in major high head stilling basins it is generally preferable to isolate the water retained in the basin
and make the basin watertight using water bars at joints. Under-drains are also generally provided and these are
best made reasonably large to help dissipate any small high head leakages and pressure surges which may enter.
On some basins drainage is only provided laterally and connected to risers in the stilling basin sidewall. This
effectively prevents the upstream-downstream connection between the different pressure zones of the basin.
Others prefer to interconnect the drains to create greater redundancy but, where this is done, sometimes the
upstream, low pressure zone is kept separate from a downstream high-pressure one. For interconnected
systems, drain exits should generally be located where external pressures are known and stable and not subject
to hydrodynamic pressure fluctuations.
Anchor bars are used for two reasons. The first is to provide yet another line of redundancy in case pressures
are transmitted below the basin. In such cases the focus is generally on reinforcing those parts of the basin
subject to reduced upper loadings against sub-pressures being transmitted from other parts of the basin where
the upper loadings are higher. Large diameter bars are generally more economic than small ones given that two
thirds of the cost of providing them is likely to be in hole drilling and installation. There are also good reasons
to stagger their lengths, with every second bar taken to just half depth. The second reason anchor bars are used
is to bond the upper slabs to the foundations thus making them less susceptible to vibration under the action of
any large, local, hydrodynamic load fluctuations within the basin. For the same reason it is generally better to
make basin apron slabs as large as practically possible to dissipate such local pressure fluctuations over larger
areas. Distribution reinforcement will clearly be needed over the upper surfaces such slabs. It is generally also
best to be generous with the concrete cover depths provided to such reinforcement. This will give some
additional protection against abrasion damage over the basin floor.
Given all the above some designers then avoid the use of shear keys between adjacent stilling basin floor slabs
so that, if one slab does fail, adjacent slabs remain unaffected. However in the writer's experience and opinion
the more common practice is to provide shear keys at joints so that any weak or susceptible floor slabs can gain
additional support from adjacent ones.

Optimum Sizing
The spillways and stilling basins for major dam and hydropower works are generally sized to be able to pass
major floods such as the Probable Maximum Flood (PMF) or the 1 in 10,000 year flood. This does not,
however, mean that the hydraulic performance needs to be optimised about such rare and infrequent events. The
hydraulic performance of stilling basins is often better optimised around a lower figure, say, the 1in 500 or 1 in
1,000 year event. It is important too to consider how all such basins and other dissipaters will operate at much
lower and more frequent flood events.
Significant economies can also be made by shortening the length of hydraulic jump stilling basins. While
standardised figures are published for the theoretical length of hydraulic jumps, generally based on the
incoming Froude number, such jumps are often quite inefficient, with much of the energy being passed
downstream. It is often acceptable to provide stilling basin with a length of approximately 60% of the

theoretical jump length, provided the works, including any downstream scour effects, are appropriately
modelled and assessed.
A downstream sill is also important on such basins as it will generate a reverse hydraulic roller immediately
downstream. This will tend to push any material excavated immediately downstream of the basin back
upstream against the sill rather than lead to a progressive deepening and eventual undermining.

Conclusions
Simple conclusions from all of the above are that good guidelines exist for the hydraulic sizing of a whole range
of alternative forms of energy dissipation devices for major spillways dams. However operational experience
has also shown many cases of failure. Oversizing the energy dissipater is not necessarily the answer in such
cases, in fact in many cases economies can safely be made by reducing the size of some such structures. The
key to successful design is to better appreciate and understand the reasons for past failures and ensure they are
not repeated. It should be remembered that simply copying past precedent is not always the safest option.
Many stilling basins are submerged and have never been dewatered and inspected. Others may have been in
existence for some years but never really tested under significant flood conditions. The broad recommended
usage ranges recommended in Figure 2 are a good starting point towards a successful and safe design. This
then needs to be coupled with the use of proven and tested, good design practices, such as those discussed in
this paper, coupled with equally good follow-up on site to ensure that those practices and details are
implemented.

Selected References and Bibliography


Chow, V.T. (1983). Open Chanel Hydraulics, McGraw-Hill, New York.
Khatsuria, R. M., (2005). Hydraulics of Spillways and Energy Dissipaters. Marcel Dekker, New York
Mason P J, "The choice of Hydraulic Energy Dissipater for Dam Outlet Works based on a Survey of Prototype
Usage", Proceedings of the Institution of Civil Engineers, Part1, (72), May 1982
Mason P J, "Practical Guidelines for the Design of Flip Buckets and Plunge Pools", International Water Power
& Dam Construction, 45 (9/10), September/October 1993
Novak, P, Moffat, A. I. B., Nalluri, C, Narayanan, R. (2001). Hydraulic Structures. Spon, London.
Thomas, H. H. (1976). The Engineering of Large Dams: Parts 1 & 2. John Wiley & Sons, London.
US Army Corps of Engineers web site (www.140.194.76.129/publications/eng-manuals)
USBR web site (www.usbr.gov/pmts/hydraulics_lab/pubs/index.cfm)
USBR (1987). Design of Small Dams. United States Department of the Interior, Bureau of Reclamation
(available on web as a free download, see above)
Zipparo, V.J., Hansen, H. (1993). Davis Handbook of Applied Hydraulics, 4th ed, McGraw Hill, New York.

The Author
Dr Peter J Mason graduated with a BSc in Civil Engineering from the Woolwich Polytechnic and with MSc and PhD
degrees in Applied Hydraulics from the City University, London. He is Technical Director, International Dams &
Hydropower for MWH Ltd with a career spanning over 43 years and over 40 countries. He has worked on major
international dam & hydropower projects in Africa, Asia, Europe and North and South America. Among other roles he
currently chairs the Board of Management for a major, 969 MW hydropower project under construction in Asia. Recent
Chairman of the British Dam Society he has also authored over 60 technical papers on all aspects of dams and hydropower.

Você também pode gostar