You are on page 1of 12

BY JOHN H.

GUMMER
Some of the most attractive hydro sites are plagued by silt at certain times of the year. Silt
erosion of the hydraulic turbines at these sites is typically controlled by upstream settling
chambers and turbine protective coatings; however, further work is needed to better predict and
control silt erosion.
The problem of hydro-abrasive erosion of hydraulic machinery is not limited to
hydroelectric plants. Similar problems to those encountered in hydroelectric work are also
prevalent in the mining industry, dredging work, and waste disposal. The blades of gas
turbines are subjected to erosion from high-velocity solid particles and those of steam turbines
from liquid droplets. Although experience gained in other industries can and is applied to similar
problems in hydraulic turbines, hydro requirements are somewhat unique in requiring high
machine efficiencies over relatively long periods between maintenance and in the
unpredictability of the concentration, shape, and mineral composition of the particle load to be
combated.
The problem of abrasive particles in hydroelectric plants is not new. European and North
American plants in alpine areas have combated the problem for years. However, the problem is
becoming more universal and more acute with the increasing need for electricity in developing
countries and the worldwide drive to promote renewable, non-fossil fuel energy sources.
Countries with the most spectacular annual increases in per capita gross domestic product,
and subsequently with the most need for economically viable, carbon-free energy, tend to
be those with the particle-laden rivers. 1 For example, many of Indias potential hydroelectric
sites are in the north of the country and are fed by run-off water from melting glaciers. Likewise,
many of South Americas viable sites are in the Andes; large rivers in China can carry substantial
quantities of particles.
Hydro-abrasive erosion of hydraulic turbines in these and other regions is an ongoing
problem and one which needs to be solved or at least mitigated if hydropower in these
countries is to reach its full carbon-free potential. The problem is exacerbated by the
favorable head and topography of many of the rivers which, at least during parts of the
year, carry high particle content. They tend to be in mountainous regions, which imply a high
head leading to a relatively low construction cost per kilowatt. They typically have fewer
environmental and resettlement problems. However, these economic advantages tend to be
somewhat negated if the development requires large expenditures on extensive settling chambers
or long periods of diminished generation due to a seasonal high particle load in the river.
MECHANISMS OF HYDRO-ABRASIVE EROSION
Tests on curved hydraulic conduits show that, on the outside curve of the conduit
(equivalent to the pressure side of a Francis turbine blade or a Pelton runner bucket), even
at relatively low velocity, particles above 1,000 micrometers (m) in diameter will not
follow the hydraulic contour, and will impact upon and damage the hydraulic
surface.2 Particles with diameters between 100 m and 1,000 m will tend to be channeled
along the outer hydraulic contour, and their propensity for damage will be progressively

less. For particle diameters below 100 m, the surface damage increases considerably. This
is because small particles become entrained in the turbulent boundary layer, which encases
all hydraulic surfaces, and results in a sand blasting of the surface.
Overall erosion from fine particles, if in sufficient quantity, can be as great as that from large
particles. Past research indicates that, due to flow separation, the inside bend surface (equivalent
to the suction side of a Francis turbine blade) experiences a steady increase in damage as the
particle size decreases. However, this generality has been contested.
The effect of particle density is similar to that of size. A particle of greater density will have
greater momentum and thus be more inclined to reach the surface in the case of larger
particles and less inclined to be entrained in the boundary layer in the case of smaller
particles. A particle can only appreciably damage a softer surface; a particle hardness of 5
Mohs is generally considered the cutoff value for hydraulic turbines. As a general rule, the
base materials used for hydraulic components in a hydro-abrasive environment should be as hard
as possible.
Theoretically, damage should be proportional to relative free stream velocity cubed. However,
probably due to boundary layer effects, this is not entirely reflected in practice. Generally
accepted experimentally determined exponents for relative free stream velocity are:

Pelton bucket, 1.5;

Pelton nozzles, 2.5;

Francis runners, 3; and

Francis wicket gates and cheekplates, 2.5.

A particles angle of attack to the hydraulic surface gives rise to two distinct erosion
mechanisms. A particle approaching nearly normal to the surface will produce impact
damage in which the surface is initially cracked, subsequently loosened with further
impacts, and finally excavated as the already cracked and loosened particle is removed by
another impacting particle. A particle approaching parallel to the surface will scratch and
gouge the surface similar to that of mechanical grinding. For angles between the two
extremes, the erosion mechanism will be a combination of both.

This hydro-abrasive particle damage on the trailing edge of a high-head Francis turbine
runner occurred after only a few months of operation in a heavily particle-laden river in
India.
The resistance to impact or gouging erosion will depend on the characteristics of the hydraulic
surface. Very hard base materials tend to be brittle and have limited resistance to impact erosion.
On the other hand, extremely hard base materials are very resistant to gouging erosion.
Conversely, soft rubbery materials (including soft plastics) exhibit a good resistance to impact
erosion but less resistance under the chiseling action of gouging erosion. These generalities do
not necessarily apply to protective coatings, for which the effectiveness of the bond between
coating and base material and the ability of both the coating and its bond to accommodate flexing
of the protected component in operation are additional, equally important, variables.
The shape of the particles is especially relevant for gouging erosion where the particle
adopts a cutting action. A sharp-edged, irregular particle can have a far greater deleterious
effect on the surface than a well-rounded or spherical particle. Under the gouging mode of
erosion, experiments indicate that the erosion rate of a jagged particle can be up to three
times that of round particle.
Particle concentration is usually expressed in grams per liter (g/l). However, often parts per
million (ppm) by weight is used, with the approximation of 1,000 ppm equal to 1 g/l being
normal usage. For a particular facility, particle concentration is essentially a temporal measure
(i.e., the greater the concentration, the higher the erosion rate and hence less time to equipment
failure). At many facilities, concentration is measured and operation of the units ceases when the
particle contraction exceeds a pre-determined value. Erosion rate is essentially proportional to
concentration over the practical operating range for a hydropower unit, but there are indications
that at higher concentrations at which hydropower stations typically no longer operate, a power
law may be more appropriate.
The effect of particles on cavitation in hydraulic reaction turbines is twofold. 3 Particles 50
m in diameter or less provide nuclei for cavitation bubbles, leading to premature
commencement of incipient cavitation at Thoma sigma values higher than in pure water.
Cavitation bubble development similarly reflects the premature commencement of cavitation.
The second, equally serious effect of hydro-abrasive erosion is that it locally changes the

hydraulic contour, which, in turn, disrupts the flow and increases both the propensity for and
intensity of the cavitation bubble implosion. The impacts from the implosion of the cavitation
bubbles fatigue and loosen the hydraulic surface, adding to the erosion damage and making
it easier for the impacting particles to remove damaged material. The combination of
cavitation and erosion is referred to as a synergistic effect, the damage resulting from
the combination of the two being far greater than the sum of each acting alone.
TURBINE COMPONENTS AFFECTED BY HYDRO-ABRASIVE EROSION
The relatively low stream velocity in the casing of a Pelton unit means that it is minimally
affected by hydro-abrasive erosion. Nozzles, however, suffer greatly; the nozzle spear and trim
erode badly, with a resulting decrease in overall turbine efficiency. The inside of Pelton buckets
suffers considerable erosion. Splitters tend to suffer impact erosion and the buckets a mixture of
impact and gouging. Turbine efficiency is compromised, especially as a result of erosion of the
splitter.
As with Pelton units, because of relatively low velocities, the scroll casing of a Francis unit is
typically immune from hydro-abrasive erosion. The nose of the stay vanes can be damaged by
impact erosion, but usually this a not a major concern. Wicket gates suffer from both impact
damage and gouging erosion. Top and bottom cover cheek plates and runner labyrinth seals
sustain considerable damage from gouging erosion.

Francis runners experience major damage at the leading edge due to impact erosion and
equally severe loss of material along the length of the blades from a combination of gouging
erosion and impact erosion. Trailing edges, because of their initial thinness, are particularly
prone to damage.
Erosion of all the above mentioned components eventually means a considerable loss of
efficiency. Losses include volumetric efficiency due to increase in labyrinth seal and guide

vane clearances, form efficiency due to the change in hydraulic profile of the wicket gates
and runner blades, and increase in frictional losses resulting from roughening of the
hydraulic surfaces. Paradoxically, the initial erosion of the trailing edges can produce an
increase in efficiency due to widening of the flow path; however, this is short-lived as the
other detrimental effects of erosion come into play. Likewise, erosion from small particles
can grind and hone the surface, leading to an initial reduction in friction. This advantage is
soon swamped by a loss in efficiency resulting from deleterious changes in the hydraulic profile.
METHODS TO PREDICT EROSION IN HYDRAULIC TURBINES
State-of-the-art computational fluid dynamics (CFD) methods are employed to further
understand the mechanics of hydro-abrasive erosion and, in particular, to design erosion-resistant
hydraulic profiles. This typically implies lower specific speeds than normal for the given
head (hence lower relative velocities), coupled with fewer jets in the case of Pelton turbines
and longer, less sharply contoured blades for Francis turbines. Where the probability of
severe hydro-abrasive erosion is high, it is sometimes economically expedient to sacrifice as
new turbine efficiency for a greater resistance to hydro-abrasive erosion. The setting of a
Francis turbine relative to minimum tailwater level should be greater than normally
accepted to ensure minimal cavitation damage.
The computerized methods successfully predict the region of maximum wear and can
somewhat mitigate the problem by refining hydraulic design and predicting the exact type
and position of protective coatings needed. However, they are incapable of accommodating all
the independent variables involved in hydro-abrasive erosion of any particular turbine. For this,
one has to resort to semi-empirical methods based upon the myriad of data from operating
turbines and from accelerated wear laboratory tests. The semi-empirical methods combine
the various factors that influence the rate of erosion under particular and definable particle
conditions but, because of the deeper setting of an erosion-prone. Francis unit, typically do
not account for the synergistic effects of cavitation and hydro-abrasive erosion.
The simplest of the various erosion criteria employs the factor H x C, where H is the net
head of the turbine in meters and C is the average annual particle concentration in g/l of all
particles with a diameter of > 50 m.4 The proposed ranges for hydro-abrasive erosion damage
risk are:

H x C = > 7: severe;

H x C = >0.7 and < 7: moderate; and

H x C = < 0.7: negligible.

Proposed by Nozaki as an extension of the Zu Yan approach is the modified particle


concentration factor, which is the product of the annual average particle concentration in

g/l and modifying coefficients related to the variables of particle size, hardness, shape, and
runner material.5 The factor is shown in Equation 1:

PE = P x a x k1 x k2 x k3

where:

PE is the modified suspended concentration in g/l;

P is the measured suspended concentration in g/l; and

Factors a, k1, k2, and k3 depend on the type and geometry of the particles and type of
runner material (see Table 1).

The final value is then used in curves of PE against turbine net head to predict times between
maintenance. Figure 1 gives an example of the curves for Francis runners. Similar charts are
available for Francis wicket gates, Pelton runners, and Pelton spears and trim.5

Figure 1 This curve shows predictions of time between maintenance for Francis runners,
based on the concentration of suspended sediment in the water and the net head of the
turbine

A more sophisticated approach, which takes into consideration the component makeup of the
particle load and accommodates variable particle concentrations in respect to time (as opposed to
average annual concentrations), is that given by the Abrasion Index. This index considers the

contribution in terms of hardness and size of each component mineral in the particle load and
then incorporates this in a formula linking the hardness of the base material and relative free
stream velocity to obtain the erosion rate.6
The main use of these empirical measures is in initial feasibility work to determine whether
hydro-abrasive erosion is going to be a problem at a site. If the answer is in the affirmative, the
project designer will want to consider incorporating particle exclusion facilities, a hydro-abrasive
erosion friendly turbine and powerhouse design, and possibly turbine protective coatings.
The various approaches to the problems of hydro-abrasive erosion typically take an overall view
and do not consider the individual effects of turbine operational efficiency at and away from the
whirl-free region in the case of reaction turbines and off-design operating conditions for
impulse units. Considering the increase in turbulence and cavitation at operation away from the
point of best efficiency, hydro-abrasive erosion will be greater the further the unit operates away
from its optimum region especially if synergistic effects are dominant.
PARTICLE EXCLUSION METHODS
The first weapon against hydro-abrasive erosion is to try to remove it before it reaches the
hydro facility. The most efficient method of removing particles is to provide a large head
water reservoir which, if of great enough volume and length, will settle out all harmful
particles. However, although very effective, it must be remembered that unless provided
with effective bottom outlets, a reservoir is only a delaying device. Eventually it will fill and,
given adverse topography, deposited particles will reach the turbine inlets.
Many mountainous hydroelectric facilities do not lend themselves to large upstream reservoirs.
In these cases, the only viable solution is settling chambers with flushing facilities. The design
of such chambers is well established, and their efficacy in removing particles greater than 1
millimeter (mm) is universally recognized. For convenience, incoming load can be divided
into three categories: coarse (>200 m); medium (75 m to 200 m); and fine (<75 m).
Reservoirs are efficient in removing all particles down to the upper fine size but settling
chambers have their limitations. Although they remove virtually all coarse particles,
depending on incoming concentration, they can pass about 80 percent of fine particles. For
medium particles, removal effectiveness is a function of particle size, falling between the two
extremes. As hard, sharp particles as small as 50 m in large quantities can cause damage
to hydraulic turbines, it is important for designers to acknowledge that, even with settling
chambers, in heavily particle-laden rivers (up to a 20 g/l peak concentration in some
Himalayan rivers7) a significant particle content will have to be accommodated in the
turbines and auxiliaries at certain times of the year.
PROTECTIVE COATINGS
As settling chambers inevitably pass substantial quantities of aggressive particles, in many
practical instances the turbine still has to be designed for, and its surfaces protected against, the
maximum expected particle load downstream of the chambers at which operation of the units is
allowed.

Two criteria for judging the effectiveness of a coating are: particle levels at which the turbine is
shut down are as high as possible; and the time between repairs is as long as possible. A third
criterion is the ease and speed by which repairs to the coating can be made.
Protective coatings fall into two categories: hard coatings such as welded Stellite and
thermally applied ceramic and tungsten carbide; and soft coatings, which are typically a brush,
trowel, or spray-on polymer. Variants of the pure hard coating are the thermally applied systems
of hard particles in a softer matrix. These hybrid systems bridge the gap between hard and soft
coatings while maintaining the potentially superior bonding strength of the thermal application
process when compared with the brush or spray-on application of soft coatings. Conversely, the
resistance of soft coatings against particle erosion depends on the type of polymer, the surface
quality, and the bond efficiency. Given the correct composition and bond for the particular
application, a soft coating can be every bit as effective against particle erosion as a hard coating.
Hard ceramic coatings have a poorer resistance to cavitation compared with soft coatings. On the
other hand, soft coatings are particularly susceptible to damage from water-borne stones and hard
debris. Soft coatings, being inherently more flexible than hard ones, can better accommodate any
movement of a protected component in service. In the case of narrow hydraulic channels, soft
coatings may be the only option owing to insufficient clearance to accommodate the equipment
for hard coating application. Hard and hybrid coatings are typically thinner than soft ones and
have more stringent thickness tolerances.
If a coating is worn in service, the old coating must be removed, the component weld repaired if
the base metal is damaged, and the coating reapplied. Being brush applied, soft coatings are
easier to repair than hard coatings. Repair of soft coatings on larger units can be done in-situ.
Hard coatings are far more difficult to repair; typically, the remaining hard coating must be
removed by grinding and a new coating applied in the suppliers factory. This is very timeconsuming but can be accelerated to a certain extent in multi-unit stations experiencing
substantial particle erosion by installing hard coating repair equipment at the project site.
The choice of a hard or soft coating depends on the characteristics of the particular component to
be protected, the hydraulic forces to which it is subjected in service, and the ease of application
and repair of the coating. In choosing the type of coating, it is prudent to remember there are
very few components that suffer only pure impact or pure gouging erosion. Typically, it is a
combination of both.
Accordingly, hard thermally applied coatings are more suited to components requiring tight
dimensional clearances such as Pelton nozzles, Pelton buckets, Francis runner wearing rings, and
Francis cheek plates. Because of their ease of repair, the prevalent use of soft coatings is on the
water passages of stay vanes (if necessary), wicket gates, and runners; however, depending on
the prevailing site conditions, hard and hybrid coatings are also used for the hydraulic surfaces of
wicket gates and runner blades.
Commercially available products can be applied to a damaged component as a mastic filler. This
application repairs the component surface and provides a resistant surface to subsequent hydroabrasive erosion. A caution: the mastic has no strength in tension; if the erosion of the base
material is substantial, the structural integrity of the component is inevitably compromised.

In general, all commercial coatings show a markedly improved hydro-abrasive erosion resistance
in service when compared with commonly used base materials such as soft Martensitic stainless
steel. In addition to actual in-service performance, coating materials are often evaluated under
laboratory conditions with the ensuing published results showing substantial lower erosion rates
than the typical base material and proprietary coatings of other suppliers. These results and those
from previous hydro projects may or may not be relevant to a particular facility. Experience has
shown that coating performance depends very much on actual site conditions. A coating may
perform well at one site or in general laboratory tests but may not be suitable for the site under
study.8 However, both laboratory testing and past experience at similar sites are good starting
points for the development of a tailor-made coating for a new development.
Contractual considerations
There is no known hydraulic machine code that specifically applies to hydro-abrasive erosion
guarantees. The International Electrotechnical Commission (IEC) Technical Committee #4s
Working Group 29 is working on this problem, and the draft code Guide for Dealing with
Abrasive Erosion in Water may be available in the future. IEC Code 60609 addresses wear of
specific turbine parts; it covers the guarantee and evaluation for cavitation damage in
reaction turbines and pumps (Part 1) and impulse turbines (Part 2). However, Clause 1.3 of
this code specifically excludes damage due to abrasive solids in the water, stating that if
relevant, the types of minerals and size of solid (sand) particles in the water analysis and, if it
reaches significant proportion, shall be the subject of a special agreement.
This statement is reflected in most equipment contracts, which give a loss of material criterion
for cavitation damage, but typically exclude erosion damage due to particles in the water. The
recommendation of IEC 60609 for the purchaser to enter into a mutually acceptable agreement
with the proposed contractor for combined cavitation and particle damage guarantee sounds
simple in principle but is extremely difficult to apply in practice. Contractors are understandably
loath to guarantee their products against particle erosion when the incoming particle conditions
cannot be entirely controlled. The possibility of synergistic cavitation and subsequent erosion
only exacerbates the situation. A cavitation guarantee is meaningless if the hydraulic surfaces are
aggressively attacked by particles.
Contractors acceptance of guarantees for hydro-abrasive erosion would be beneficial to both
purchaser (in order for him to assure unit output) and the contractor, in order for him to receive
commensurate consideration during bid evaluation of his research and development expenditure
on particle damage.
A major obstacle to obtaining guarantees for particle damage from a manufacturer is the problem
of exactly defining the particle load and monitoring it during the guarantee period to ensure
compliance with the limits stipulated in the contract. Historical records are obviously a guide to
establishing the contractual particle load, but, because of the unknowns of the efficacy of
settlement in reservoirs, power channels, and settling chambers, these may not be adequate for
guarantee purposes. In addition, a change to the historical particle load is possible because of
upstream landslides, an event which is almost impossible to design for.

The current methods for monitoring particle load are sample taking and off-line analysis, which
is slow and labor intensive. On-line optical or acoustic diffusion techniques are being
developed.9 Coupled with the problems of defining and monitoring the actual particle load is the
possibility of gross damage to any protective surface by large debris (stones, rocks, and hard
foreign objects); however, this is unlikely (but not impossible) if settling facilities are installed
and the inlet water channels are lined.
Further research and development into all aspects of combating hydro-abrasive erosion is
necessary if the protection of hydraulic turbines against this form of damage is to be put
onto a contractual basis as covered by an appropriate international code acceptable to all
parties.

NOTES
1. Krishnamachar, P., and S. Rangnekar, Correlation of Hydropower Potential of Silt Load
of Rivers Means to Access Damage by Silt Due To Not Harnessing Hydropower,
HydroVision 2008 Conference Papers CD-Rom, HCI Publications, Kansas City, Mo.,
USA, 2008.
2. Ortmanns, C., and S. Prigent, Turbine Abrasion and Desilting Chamber Design, Hydro
2006 Conference Papers CD-Rom, Aqua-Media International. Ltd., Sutton, Surrey,
United Kingdom, 2006.
3. Shengcai, L.L., Cavitation Enhancement in Silt Erosion: Obstacles and Way Forward,
5th International Symposium of Cavitation, Osaka, Japan, 2003.
4. Zu-Yan, M., Review of Research on Abrasion and Cavitation of Silt Laden Flows
through Hydraulic Turbines in China, 18th IAHR Symposium on Hydraulic Machinery
and Cavitation, International Association of Hydraulic Engineering and Research,
Valencia, Spain, 1996.
5. Nozaki, T., Technical Report: Estimation of Repair Cycle of Turbine Due To Abrasion
Caused By Suspended Sediment and Determination of Desilting Basin Capacity, Japan
International Cooperation Agency, 1990.
6. Sharma, S.K., Sediment Management in Himalayan Rivers, HydroVision 2006
Conference Papers CD-Rom, HCI Publications, Kansas City, Mo., USA, 2006.

7. Dhar, D.K., and P. Dul, Silting Problems in Hydropower Plants, Waterpower XV


Conference Papers CD-Rom, HCI Publications, Kansas City, Mo., USA, 2007.
8. Sharma, M.K., G.S. Grewal, and A.K. Singh, Silt Erosion in Indian Hydroelectric
Projects Laboratory Studies of Thermal Spray Coatings over Hydro Turbine
Components, HydroVision 2008 Conference Papers CD-Rom, HCI Publications, Kansas
City, Mo., USA, 2008.
9. Bishakarma, M.B., Online Monitoring of Sediments in Hydropower Plants: A System
for Assessing the Turbine Exposure and Sediment-Induced Effects, Waterpower XIV
Conference Papers CD-Rom, HCI Publications, Kansas City, Mo., USA, 2005.
Mr. Gummer is a director of Hydro-Consult Pty Ltd., a company specializing in feasibility
studies, specification, and contract administration of mechanical and electrical equipment
for hydroelectric projects. John has worked for 45 years on equipment specification and
equipment contractual problems with the World Bank and about 20 government
organizations and private consulting companies on numerous hydro projects, many of
which have been in hydro-abrasive prone environments.

MAINTENANCE-FRIENDLY DESIGNS
A unit operating with particle-laden water should be designed for ease of maintenance and repair.
This makes the standard barrel-type vertical design for Francis turbines, where the generator
must be removed in order to access the turbine parts, unsuitable other than for very large units
where repair in situ is possible.
Small- to medium-sized vertical Francis units typical in high to medium head plants
should be designed for bottom dismantling of the runner, bottom ring, discharge ring, and wicket
gates. This design provides for access to these components and the head cover and check plates
for repair. Because of demand for this feature, the size of Francis turbines with this arrangement
is increasing. Runner diameters of up to 5.1 meters have been constructed with bottom
dismantling.
The question of whether to choose Pelton or Francis turbines has been ongoing for many years.
There is no clearcut answer, as both suffer from hydro-abrasive erosion. A Pelton unit is easier to
maintain, but civil costs will be higher in order to accommodate the larger machine. This may be
compensated, in part, in a surface facility by the cost of the additional excavation required for
Francis unit. Each project has to be considered separately and a reasoned judgment made based
on economics, ease of maintenance, site topography, and geology.

Particles can erode heat exchanger tubes if velocities are too high and block them if velocities are
too low. If particles are known to be a problem, then closed loop cooling water systems are
installed with duplicate raw water/closed loop water heat exchangers. Raw water piping should
be stainless steel. The turbine shaft seal should be designed for hydro-abrasive conditions and, if
required by the design, provided with a supply of well-filtered water to reduce wear.
Availability of spare runners, nozzles, wicket gates, and cheek plates enable damaged parts to be
quickly replaced and subsequently repaired and/or recoated for re-installation during the next
maintenance outage.

You might also like