You are on page 1of 8

HYDRAULICS OF SKIMMING FLOW ON teristic flow depth for the air-water mixture, which would cor-

respond to d10 according to the authors’ definition—i.e., the


MODELED STEPPED SPILLWAYSa probe is immersed in water 10% of the total measuring time.
Another physically meaningful flow depth in two-phase flow
is the clear water depth dw = d10(1 ⫺ C̄), because due to the
low density of the gas phase of about 1/800 of pure water,
Discussion by Robert M. Boes4 aeration can be neglected in the calculation of Reynolds shear
stresses. The depth-averaged air concentration C̄ is deduced
by integrating C-profiles up to d = d10 = d(C = 0.9). Why did
INTRODUCTION
the authors choose d50 as the characteristic flow depth?
The authors presented an interesting model study on two-
phase cascade flow using different step heights in a large EFFECT OF STEP HEIGHT H
model flume. Their results underline the high energy dissipa-
Similar to the authors’ findings on y2 being independent of
tion capacity of stepped spillways, allowing a considerable re-
the step height H, the discusser found no influence of H on
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

duction of the downstream stilling basin and thus making them


the uniform clear water depth hw,u perpendicular to the spillway
attractive to dam designers.
chute for his data. The parameter hw,u is constant for a given
The discusser conducted model tests on skimming flow in
discharge and slope and can be fitted by a trendline, as shown
a large model of a stepped spillway for different step heights
in Fig. 14.
and two slopes, 1.73:1 (30⬚) and 0.84:1 (50⬚), with specific
prototype discharges up to 50 m2/s. The steeper chute was only
EQUILIBRIUM AS FUNCTION OF HEIGHT
slightly outside the 0.6–0.8 slope range of ‘‘comparable steep-
BELOW CREST
ness’’ according to the authors. The following remarks and
questions are raised: The discusser’s data also confirm the authors’ assumption
of increasing mean velocity of the two-phase flow with de-
RESIDUAL SPECIFIC FORCE AND SPECIFIC ENERGY
The discusser would like to draw the authors’ attention to
measuring devices such as electrical resistivity (Cain and
Wood 1981; Volkart 1988; Chanson 1993; Frizell et al. 1994;
Ruff and Frizell 1994; Matos and Frizell 1997) or fiber-optical
probes (Boes and Hager 1998; Boes, unpublished, 2000),
which allow direct measurements of velocity and air concen-
tration distributions in highly turbulent two-phase flow. The
knowledge of the depth-averaged air concentration C̄ permits
calculation of the theoretical clear water depth dw that should
be used to compute the momentum flux and hence the specific
energy at the toe of a spillway.

COMPARISON OF RESIDUAL SPECIFIC ENERGIES


Given the steep chute studied by the authors, it is likely that
the specific energy E1 on a horizontal basin is significantly
smaller than the specific energy Eslope on the sloping part at
the end of the chute, because the additional head loss resulting
from the jet deflection on the horizontal basin increases with FIG. 14. Uniform Clear Water Depth hw,u versus hc /(sin ␾)1/3 in
increasing inclination angle ␾. Therefore, to be on the safe Prototype Dimensions for 1:19.6 Scale Model with Two Different
side in the design of a stilling basin, Eslope should be accounted Step Heights and 0.84:1 Slope
for rather than E1 as long as the additional energy loss cannot
be reliably quantified.

APPARATUS AND METHOD OF MEASUREMENT


The authors used an electronic contact probe to measure the
so-called connected surface, i.e., the flow depth separating the
two typical flow zones in free-surface two-phase flow. The
lower zone contains air bubbles carried by the water flow,
whereas the upper zone comprises water droplets in air. The
authors defined the flow depth d50 with equal proportion of
time for the water and the gas phases as the characteristic flow
depth perpendicular to the flow direction. Thus, for steady
flow, d50 is the depth with a local air concentration C of 50%,
i.e., d50 = d(C = 0.5), where C is defined as the volume of air
per unit volume of air and water, or the proportion of time the
probe is in gas. It should be borne in mind, however, that a
flow depth with C = 0.9 is commonly considered the charac-
a
May 1999, Vol. 125, No. 5, by Geoffrey G. S. Pegram, Andrew K. FIG. 15. Characteristic Flow Depths and Mean Air-Water Ve-
Officer, and Samuel R. Mottram (Paper 3557). locity versus Distance from Inception Point x-xi in Model Dimen-
4
PhD Student, Lab. of Hydr., Hydrol., and Glaciology (VAW), Federal sions for 1:19.6 Scale Model with H = 31.1 mm, hc /H = 2.44, and
Inst. of Technol. (ETH), CH-8092 Zurich, Switzerland. 0.84:1 Slope

JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000 / 947

J. Hydraul. Eng. 2000.126:947-954.


creasing median depth d50. Fig. 15 shows characteristic flow EFFECT OF STEP HEIGHT ON ENERGY DISSIPATION
depths as well as the depth-averaged velocity ūm of the air-
One major conclusion derived from the authors’ study was
water mixture in terms of the distance from the inception point
that the energy dissipation ratio (EDR) is independent of the
(subscript i) x ⫺ xi for a typical experimental run. It is ob-
step height H in the range of flows and step heights tested.
served that (1) the median flow depth d50 is within 20% of the
This conclusion, of great importance for the design of stepped
clear water depth dw, (2) both d50 and dw follow drawdown
spillways, seem to strengthen the concept of the ‘‘optimum
curves, and (3) the characteristic mixture flow depth d10 rep-
step height,’’ as suggested by Matos and Quintela (1995a) after
resents a backwater curve. The mean velocity ūm was deduced
the findings of Stephenson (1991) and Tozzi (1992).
from the velocity profiles by integration up to d = d10.
Stephenson (1991) suggested that the energy dissipation can
be increased until the stage when the water depth is approxi-
INFLUENCE OF SCALE mately one-third of the critical depth (d/hc = 1/3), beyond
For both slopes examined in the discusser’s model studies, which it is very difficult to increase the energy dissipation any
the results on air concentrations and flow velocities of the air- further. A close value of d/hc (d/dhc = 0.294) was proposed by
water mixture were compared for different scales. Taking a Tozzi (1992), based on data gathered on a 53⬚ stepped chute,
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

common prototype step height of H = 0.61 m as reference for 0.125 ⱕ H (m) ⱕ 1.50 and 1.37 ⱕ hc (m) ⱕ 2.40 (values
length, practically no scale effects occurred for model scales scaled to prototype). Tozzi (1992) has also suggested that the
larger than 1:13.2, and only minor scale effects were observed maximum roughness height k (k = H cos ␣) is given by k =
for the 1:19.6 and 1:26.4 scales, which is in close agreement 0.0764 q2/3, hence H/hc = 0.27. In Fig. 16, Tozzi’s data re-
with the authors’ findings. More details on scale effects are garding the ratio of the energy dissipated to the total energy
given by Boes (unpublished, 2000). is plotted as a function of H/hc. In light of results obtained by
Tozzi (1992) and by Diez-Cascon et al. (1991), Matos and
APPENDIX. REFERENCES Quintela (1995a) suggested that the step height above which
the increase in the energy dissipation is negligible (optimum
Boes, R. M., and Hager, W. H. (1998). ‘‘Fiber-optical experimentation in step height) can be estimated by Hopt /hc ⬇ 0.3. This relation
two-phase cascade flow.’’ Proc., Int. RCC Dams Seminar, K. Hansen, is also represented in Fig. 16, where it can be seen that for
ed., Denver. H/hc ⱖ 0.3, the energy loss ratio, (Em ⫺ Et)/Et, remains prac-
Cain, P., and Wood, I. R. (1981). ‘‘Instrumentation for aerated flow on
spillways.’’ J. Hydr. Div., ASCE, 107(11), 1407–1423. tically constant (its maximum relative increase was found to
Chanson, H. (1993). ‘‘Velocity measurements within high velocity air- be less than 3%).
water jets.’’ J. Hydr. Res., Delft, The Netherlands, 31(3), 365–382. The observation of Fig. 3 shows that the energy loss is in-
Frizell, K. H., Ehler, D. G., and Mefford, B. W. (1994). Developing air dependent of the step height H, for H greater than 0.25 m and
concentration and velocity probes for measuring highly-aerated, high- hc less than about 1 m (i.e., H/hc > 0.25). In Fig. 5, an identical
velocity flow. Proc., ASCE Conf. Fundamentals and Advancements in conclusion is obtained for H greater than 0.5 m and hc less
Hydr. Measurements and Experimentation, C. A. Pugh, ed., American
Society of Civil Engineers, New York, 268–277. than about 2.3 m (i.e., H/hc > 0.22). Therefore, the authors’
Matos, J., and Frizell, K. H. (1997). ‘‘Air concentration measurements in relevant findings are in agreement with those based on the data
highly turbulent aerated flow.’’ Proc., 27 IAHR Congr., S. S. Y. Wang of Tozzi (1992) and of Diez-Cascon et al. (1991).
and T. Carstens, eds., A: International Association of Hydraulics Re- It is interesting to note that the step heights widely used on
search, Delft, The Netherlands, 149–154. RCC dams, 0.60 m and 0.90 m (or 0.5 and 1.0 m, for RCC
Ruff, J. F., and Frizell, K. H. (1994). ‘‘Air concentration measurements layers of 0.25 m), are not too far from the most efficient ones,
in highly-turbulent flow on a steeply-sloping chute.’’ Proc., ASCE Nat.
Conf. on Hydr. Engrg., G. V. Cotroneo and R. R. Rumer, eds., ASCE, for unit discharges ranging from 10 to 20 m2/s. Considering,
New York, 2, 999–1003. however, that higher RCC dams are currently being proposed,
Volkart, P. (1988). ‘‘Instrumentation for measuring local air concentration the validation (or reformulation) of the relation Hopt /hc = 0.3
in high-velocity free-surface flow.’’ Proc., Int. Symp. on Hydr. for High is judged of interest, on the basis of experimental tests on
Dams, Beijing, China, 1088–1096. higher stepped chutes and for higher unit discharges.

SELF-AERATED SKIMMING FLOW AT EQUILIBRIUM

Discussion by Jorge Matos5 By lifting the hydraulic jump basin to two levels above the
floor at the toe of a 1:20 scale model of a 58-m-high stepped
spillway, the authors found that equilibrium as measured by
INTRODUCTION y2 appeared to have been obtained at or before Y = 50 m

In their important contribution, the authors draw attention


to topics of major interest for the hydraulic design of stepped
spillways over RCC dams, such as the effect of step height on
energy dissipation, equilibrium as a function of height below
crest, and scale effects. Innovative findings of skimming flow
over stepped spillways have also been included, namely, the
wavy pattern of the profiles of the median flow surface.
In the present discussion, some of the authors’ relevant re-
sults are analyzed in light of other information available in the
literature, namely, on the effect of step height on energy dis-
sipation (Diez-Cascon et al. 1991; Tozzi 1992), air entrainment
and equilibrium (e.g., Lejeune et al. 1994; Matos and Quintela
1995a; Chamani and Rajaratnam 1999), waviness of the char-
acteristic depth profiles (Matos et al. 1999), and scale effects
(Diez-Cascon et al. 1991; Tozzi 1992; Boes and Hager 1998).
FIG. 16. Effect of Step Height on Energy Dissipation—Opti-
5
Lect., Technical Univ. of Lisbon, IST, Dept. of Civ. Engrg., Lisbon mum Step Height [Experimental Data by Tozzi 1992: 0.125 ⱕ H
1096, Portugal. (m) ⱕ 1.50, at 1:15 Scale]

948 / JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000

J. Hydraul. Eng. 2000.126:947-954.


authors for the median depth profiles (due to both the en-
trapped and entrained air, the median depth is expected to lie
between d and Y90). Another interesting result of Fig. 18 is that
the mean depth obtained by visual observation dmean is similar
to Y90 (sufficiently downstream of the inception point) and
much greater than the equivalent clear water depth, d. This
confirms the significant overestimation of the energy loss that
can occur when the bulked flow depth is used to compute the
specific energy, as suggested in previous studies (e.g., Matos
and Quintela 1995a,b).

SCALE EFFECTS
On the basis of experimental data gathered on 1:10, 1:20
(authors’ data), and 1:15 (Tozzi’s data) modeled stepped spill-
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

ways, the authors have concluded that models with scales of


1:20 and larger seem to faithfully represent the prototype be-
havior of stepped spillways. It was also shown that results for
FIG. 17. Mean Air Concentration in Skimming Flow over Steep stepped spillway models appear to be converging rather
Stepped Chutes (after Matos and Quintela 1995a) quickly as the scale gets bigger than 1:15.
The results of the recent study of Boes and Hager (1998),
(vertically) below the crest, within the range of flows corre- in which air concentration and velocity profiles were collected
sponding to hc in the range 0.8–2.0 m (i.e., 25.0 ⱕ Y/hc ⱕ in the self-aerated skimming flow, were found to be in agree-
62.5). As also mentioned by the authors, the median depth ment with the authors’ relevant conclusion. In their study,
profiles (Fig. 9) show that the depths, although fluctuating, are model scales of 1:6.6, 1:13.2, and 1:26.4 were used with ref-
tending to reasonably constant values before Y = 50 m, at least erence to a prototype height of 0.61 m. Therein it can be ob-
for the lower flows. This observation is patent in Fig. 9, served that similitude between the 1:6.6 1:13.2 scales is good
namely for the profiles where hc = 1.14, 1.43, and 1.73 m (i.e., for both the air concentration and velocity distributions. It
Y/hc = 43.9, 35.0, and 28.9). The authors’ findings are in ac-
cordance with the results expressed in Fig. 17, where the
depth-averaged mean air concentration near the toe of stepped
spillways is expressed as a function of Y/hc. Also included in
the same figure is the experimental data recently presented by
Chamani and Rajaratnam (1999) on stepped chutes of slopes
1V:0.8H and 1V:0.6H, the latter being equal to that of the
authors. From Fig. 17 it can be concluded that a tendency of
increase of Cmean with Y/hc is observed. For Y/hc greater than
about 30, the rate of growth of Cmean is small and it seems
reasonable to consider that the flow is quasi-uniform, similar
to the authors’ findings. However, it appears that equilibrium
condition may not go to hc = 3 m or beyond, for a 50 m high
spillway (i.e., Y/hc ⱕ 16.7). This reasoning seems to be sup-
ported by a recent analysis of air entrainment down a 1V:
0.75H chute slope (Matos and Frizell 1997). Although the air
concentration profiles suggest that a quasi-equilibrium condi-
tion was verified at Y/hc = 21, subsequent results have con-
firmed that the uniform flow was not attained for higher dis-
charges (e.g., Matos et al. 1999). FIG. 18. Characteristic Depth Profiles in Skimming Flow on
It should be noted that most of the data plotted in Fig. 17 1V:0.75H Stepped Chute: h = 8 cm, hc = 0.1 m (after Matos et al.
refers to chute slopes of about 1V:0.75H. However, signifi- 1999)
cantly dissimilar conclusions are not expected for chute slopes
of 1V:0.60H. In actual fact, the data of Chamani and Rajar-
atnam (1999) show that Cmean is practically independent of the
chute slope (ranging from 1V:0.8H to 1V:0.6H).

CHARACTERISTIC DEPTH PROFILES


One of the interesting new findings of the authors is the
wavy pattern of the profiles of the median flow surface. A
similar wavy pattern was also observed on the characteristic
depth profiles Y90, where Y90 is defined as the depth where the
local air concentration is 90% (Fig. 18). In Fig. 18 it can be
noted that the flow depths obtained by visual observation
through the sidewalls (dmin, dmax, and dmean) also exhibit a wavy
pattern. On the other hand, the waviness is not evident on the
profiles of the equivalent clear water depth, d[d = (1 ⫺
Cmean)Y90]. From Fig. 18 it can also be noted that on the dmean
and Y90 profiles, the amplitude of the waveform appears to die FIG. 19. Froude Number at Downstream End of Hydraulic
out down the stepped chute, similar to the observation of the Jump Formed at Toe of Stepped Chutes

JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000 / 949

J. Hydraul. Eng. 2000.126:947-954.


can also be observed that similitude between different scales Discussion by I. Ohtsu,6 Y. Yasuda,7
(1:6.6, 1:13.2, and 1:26.4) is acceptable for the velocity dis-
tributions. and M. Takahashi8
A comparison of the data of Diez-Cascon et al. (1991)
(1:10 scale) and Tozzi (1992) (1:15 scale) with that of the The authors examined the subcritical depth of the jump im-
authors also provides some interesting results. In Fig. 19, the mediately below stepped-channel models in order to study the
Froude number at the downstream end of the hydraulic jump equilibrium condition. Also, the residual energy at the stepped
(F2) is plotted as a function of the critical depth (hc). The channel end was shown. Recently, the discussers have inves-
experimental results of Diez-Cascon et al. (1991) and Tozzi tigated the equilibrium condition for skimming flows by con-
(1992) suggest that F2 might be considered independent of the sidering the jump formation immediately below the stepped
unit discharge and step height. However, the values are con- chute (Yasuda and Ohtsu 1999). The discussers would like to
siderably different: for the data of Diez-Cascon et al. (1991), comment on the characteristic of the jump immediately below
F2 is almost equal to 0.2, whereas for the data of Tozzi (1992), the stepped chute under the equilibrium condition.
a value of 0.25 was obtained. The expressions presented by
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

the authors in Fig. 10, for the 1:10 and 1:20 model studies, JUMP FORMATION IMMEDIATELY BELOW
can be rewritten as STEPPED CHUTE
F2 = 0.266h 0.026
c (1:20 scale) (3) Regarding the subcritical depth of the jump formed imme-
F2 = 0.218h 0.170
c (1:10 scale) (4) diately below the end of a stepped channel, the following func-
tional relation is obtained from dimensional considerations:

冉 冊
In conformity with the results from the other mentioned
studies, Eqs. (3) and (4) indicate that the unit discharge (or hc) y2 Hdam H
has not a significant influence on F2. Within the range of dis- =f , , tan ␪ (5)
hc hc hc
charges analyzed by the authors, F2 varies approximately be-
tween 0.26 and 0.27 (1:20 scale) and between 0.19 and 0.22 6
(1:10 scale). The latter values lie close to that obtained by Prof., Dept. of Civ. Engrg., Nihon Univ., Coll. of Sci. and Technol.,
Kanda Surugadai 1-8 Chiyoda-ku, Tokyo 101-8308, Japan.
Diez-Cascon et al. (1991), also at a 1:10 scale, and the former 7
Asst. Prof., Dept. of Civ. Engrg., Nihon Univ., Coll. of Sci. and Tech-
ones are not significantly dissimilar to those estimated from nol., Tokyo, Japan.
the data of Tozzi (1992), at a 1:15 scale. It is worth noting 8
Res. Assoc., Dept. of Civ. Engrg., Nihon Univ., Coll. of Sci. and
that in the authors’ 1:10 model study, equilibrium was not Technol., Tokyo, Japan.
reached at the toe of the spillway, whereas in most of the tests
of Diez-Cascon et al. (1991), quasi-uniform flow seem to have
been reached (Fig. 17). Hence, slightly higher values of F2
would be expected from the Diez-Cascon et al. (1991) study,
in comparison with those obtained by Eq. (4). A possible rea-
son for the above result might be related to the dissimilar
position of the jump adopted on the tests of Diez-Cascon et
al. (1991). In fact, if the toe of the spillway is drowned by the
jump even slightly, y2 will likely be overestimated (i.e., F2 will
be underestimated).

APPENDIX. REFERENCES
Boes, R. M., and Hager, W. H. (1998). ‘‘Fiber-optical experimentation in FIG. 20. Discussers’ Definition of Jump Position
two-phase cascade flow.’’ Proc., Int. RCC Dams Seminar, K. Hansen,
ed., EUA, Denver.
Chamani, M. R., and Rajaratnam, N. (1999). ‘‘Characteristics of skim-
ming flow over stepped spillways.’’ J. Hydr. Engrg., ASCE, 125(5),
500–510.
Diez-Cascon, J., Blanco, J. L., Revilla, J., and Garcia, R. (1991). ‘‘Studies
on the hydraulic behaviour of stepped spillways.’’ Water Power & Dam
Constr., Sept., 22–26.
Lejeune, A., Lejeune, M., and Lacroix, F. (1994). ‘‘Study of skimming
flow over stepped spillways.’’ Proc., Int. Conf. on Modelling, Testing
and Monitoring for Hydro Powerplants, HP&D, Budapest, Hungary,
July, 285–294.
Matos, J., and Frizell, K. H. (1997). ‘‘Air concentration measurements in
highly turbulent aerated flow.’’ Proc., 28th IAHR Congr., Theme D,
Vol. 1, Sam S. Y. Wang and Torkild Carstens, eds., International As-
sociation for Hydraulics Research, Delft, The Netherlands, 149–154.
Matos, J., and Quintela, A. (1995a). ‘‘Flow resistance and energy dissi-
pation in skimming flow over stepped spillways.’’ Proc., 1st Int. Conf.
on Water Resour. Engrg., ASCE, New York, Vol. 2, 1121–1126.
Matos, J., and Quintela, A. (1995b). ‘‘Energy dissipation in skimming
flow over stepped spillways. A comparative analysis.’’ Proc., 26th
IAHR Congr., London, Vol. 1, 370–372.
Matos, J., Sánchez, M., Quintela, A., and Dolz, J. (1999). ‘‘Characteristic
depth and pressure profiles in skimming flow over stepped spillways.’’
Proc., 28th IAHR Congr. (CD-ROM), Theme B, Graz, Austria.
Stephenson, D. (1991). ‘‘Energy dissipation down stepped spillways.’’
Water Power & Dam Constr., Sept., 27–30.
˜
Tozzi, M. J. (1992). ‘‘Caracterizaçao/comportamento de escoamentos em
vertedouros com paramento em degraus (Hydraulics of stepped spill- FIG. 21. Relative Downstream Depth of Hydraulic Jump (Jump
ways).’’ PhD thesis, University of Sao ˜ Paulo, Brazil (in Portuguese). Location Is Determined according to Discussers’ Definition)

950 / JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000

J. Hydraul. Eng. 2000.126:947-954.


Here, hc = critical depth; Hdam = height of dam; H = step
height; and ␪ = channel slope.
In the discussers’ experiment, the toe of the jump is defined
as the section where the bed-pressure is at its maximum, due
to the curvature of the streamline (Fig. 20).
Fig. 21 shows an example of the relation of (5) for ␪ = 55⬚
and 0.6 ⬉ H/hc ⬉ 1.25. As shown in Fig. 21, the subcritical
depth of the jump y2/hc changes with the relative dam height
Hdam/hc and is independent of the relative step height H/hc. If
the value of the relative dam height Hdam/hc is larger than 28,
the value of y2/hc becomes constant (i.e., y2/hc = 2.55) and the
equilibrium condition is obtained for given ␪ and H/hc (Fig.
22).
Under the equilibrium condition, the velocity of a quasi-
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

uniform flow can be expressed as (6).

V= 冑 8g d sin ␪
f
兹 (6)

where d = supercritical depth above the pseudobottom of a


quasi-uniform flow and f is the friction factor defined as FIG. 24. Relative Downstream Depth of Hydraulic Jump (Com-
parison of Authors’ Data with Discussers’ Data; Case A—Toe
4␶0 8gd sin ␪ Location of Jump in Fig. 20; Case B—Toe Location of Jump in
f= = (7) Fig. 23)
1 V2
␳V 2
2
y2 /d1 depends on the Froude number F and the channel slope
By using (6), the supercritical Froude number of a quasi- ␪ (Ohtsu and Yasuda 1991):
uniform flow is also expressed as

F=
V
兹gd cos ␪
= 冑8
f
兹tan ␪ (8)
y2
d1
= f(F, ␪) (9)

If the jump is formed immediately below a sloping chute, Also, the depth ratio y2/hc is expressed as
the momentum equation shows that the sequent depth ratio
y2 y2/d1 y2/d1
= = (10)
hc hc /d1 F2/3 cos1/3␪

Accordingly, substituting (8) and (9) into (10), the following


relation is obtained:

y2
= f(f, ␪) (11)
hc

It was confirmed experimentally that the friction factor f is


almost constant within the range of 0.6 ⬉ H/hc ⬉ 1.25 for ␪
= 55⬚ (Yasuda and Ohtsu 1999).
Thus

y2
= const [y2 = (const) ⫻ hc] (12)
hc

If the toe of the jump is located at the vertical step face of


the stepped chute end (Fig. 23), the value of the subcritical
depth of jump y2/hc is always larger than that in which the toe
location is determined according to the discussers’ definition
FIG. 22. Jump Formation under Equilibrium Condition (Fig. 24). Also, as shown in Fig. 24, the authors’ data coincides
with the discussers’ data of y2/hc in which the toe of the jump
is located at the vertical step face of the stepped chute end.

APPENDIX. REFERENCES

Ohtsu, I., and Yasuda, Y. (1997). ‘‘Characteristics of flow conditions on


stepped channels.’’ Proc., Theme D, 27th Congr. of IAHR, San Fran-
cisco, IAHR, Delft, The Netherlands, 583–588.
Ohtsu, I., and Yasuda, Y. (1991). ‘‘Hydraulic jump in sloping channels.’’
J. Hydr. Engrg., ASCE, 117(7), 905–921.
Yasuda, Y., and Ohtsu, I. (1999). ‘‘Flow resistance of skimming flow in
FIG. 23. Toe Section of Jump Located at Vertical Step Face of stepped channels.’’ Proc., 28th Congr. of IAHR, Theme B, Hyr. Struc-
Stepped Chute End ture, Spillways and Chute Struct., B14, IAHR, Delft, The Netherlands.

JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000 / 951

J. Hydraul. Eng. 2000.126:947-954.


Discussion by S. P. Tatewar,9 and R. N. Ingle,10
and P. D. Porey11

The authors are to be commended for publishing extensive


experimental data of skimming flow on stepped spillway. The
author’s data predicted that residual specific energy at the toe
of spillway is independent of step height, as shown in Figs.
11 and 12. At the same time, the author’s data shown in Fig.
13 indicates that the variation of EDR with hc is different for
different step heights, which appears to be contradicting the
conclusion that EDR is independent of step height H.
The discussers are of the opinion that increase in step height
increases the roughness of bed surface and therefore should FIG. 26. Notations for Dimension of Step
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

increase energy loss over the surface of spillway. The energy


loss over stepped spillway surface would depend on the ratio height in the case of Tozzi’s data as well as the Yildiz and Kas
hc /H, Froude number of flow, and average slope of spillway data. Though the variation is of the order of 10–20%, the trend
surface. Therefore, for a given slope of spillway surface it of variation indicates that with the increase in step height en-
would depend on hc /H and the Froude number of flow. The ergy loss over the stepped spillway increases and, therefore,
author’s observations in Fig. 12 are for the range of hc /H from available specific energy at the toe decreases.
0.1 to 7 and Froude number at the toe from 7 to 15, and The discussers (1996) have derived the expression for com-
probably for this range the effect of step height becomes neg- putation of equivalent Manning’s coefficient of roughness (n)
ligible. For a higher Froude number, a circular vortex is for stepped spillway in terms of size of step from which energy
formed in the gap of steps and it functions as a roller over loss in skimming flow over stepped spillway can be computed.
which flow occurs; therefore, the effect of increasing step The expression derived by the discussers is given below:

冉冊 冉 冊
height as negligible for the author’s range of experimental
data. However, for lower values of Froude number the effect Z 0.1 ␭ Z 0.6
= 0.25 ⫹ 19 log ⫹ 5.75 log (13)
of step height on energy dissipation should be considerable. n兹g b k
The effect of step height on the energy available at the toe of
spillway (E1) based on Tozzi’s (1994) data as well as on Yildiz where the variables defined with respect to Fig. 26 are Z =
and Kas (1998) data, converted to prototype value using 1:20 nq/兹S = section factor for computation of uniform flow; q =
scale, is shown in Fig. 25 along with the experimental data of discharge per unit width; b = tread length of step (equal to
the author give in Fig. 12. From Fig. 25 it is seen that the length of step); k = height of roughness; ␭ = roughness spac-
residual specific energy available at the toe varies with step ing; and S = average slope of spillway.
Values of E1 computed from (13) for different values of y2
9
Asst. Prof., Dept. of Civ. Engrg., Government Coll. of Engrg., Am- corresponding to the author’s data are plotted in Fig. 25, which
ravati – 444 604, India. show the effect of step height and good agreement with the
10
Prin., Yeshwantrao Chavan Coll. of Engrg., Wanadongri, Nagpur, Yildiz and Kas data as well as with Tozzi’s data. However,
India. these lines converge for the author’s data at higher values
11
Prof., Dept. of Civ. Engrg., Visvesvaraya Regional Coll. of Engrg.,
Nagpur – 440 010, India.
of y2.

APPENDIX. REFERENCES
Tatewar, S. P., and Ingle, R. N. (1996). ‘‘Resistance to skimming flow
over stepped spillway.’’ Proc., Int. Seminar on Civ. Engrg: Practices
in Twenty-first Century. Roorkee, India, 1039–1048.
Yildiz, D., and Kas, I. (1998). ‘‘Hydraulic performance of stepped chute
spillway.’’ Hydropower and Dams (4), 64–70.

Closure by Geoffrey G. S. Pegram,12


Andrew K. Officer,13 and
Samuel R. Mottram14

The writers were delighted by the response of the discussers


to their paper. In particular, much satisfaction was derived
from the reporting of recent results that corroborated the (in
some cases) tentative findings reported in the paper. This is
particularly pleasing because the paper was seven years in re-
view and rewrite from its original submission in February
1992, gaining much from the patient input of the reviewers
12
Prof. of Hydr. Engrg., Dept. of Civ. Engrg., Univ. of Natal, Durban,
4041, South Africa. E-mail: pegram@eng.und.ac.za
13
Hydr. Engr., Keeve Steyn Incorporated, Pietermaritzburg, South Af-
rica.
14
FIG. 25. Comparison of (13) with Authors’, Tozzi’s, and Yildiz Hydr. Engr., Ninham Shand Incorporated, Pietermaritzburg, South
and Kas’ Experimental Data Africa.

952 / JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000

J. Hydraul. Eng. 2000.126:947-954.


over that time, when originally the only results available for momentum at the base of the spillway than on the residual
comparison were those of Diez-Cascon et al. (1991). The writ- specific energy favored by others. We therefore cannot agree
ers feel that the discussion is particularly valuable because of with Boes that Eslope be the variable of choice; rather we
the enrichment of the study by the additional references and suggest that y2/hc values should be found as a function of
insight provided. dam height Ymax and slope angle ␾ for given ranges of step-
Matos takes pains to demonstrate corroboration between re- height H.
cent work and ours. The particular points of congruence are Boes observes that d50 is the depth where local air concen-
in the paper’s conclusions that, ‘‘in the range of flows and tration is 50%. This is correct if d is given as a function of
step-heights tested,’’ the EDR is independent of step height H, concentration, but not if d is a function of relative immersion
equilibrium is achieved by 50 m (making discussions of EDR as derived from the contact probe used in the study. The reason
meaningful), and scale-effects reduce rapidly from 1:20 and for choosing the d50 value to give the characteristic depth of
larger (which was only a tentative suggestion in the paper). the flow profiles (and in calculating y2) is that it was the most
Matos’ discussion on his Fig. 17 suggests that there is an repeatable measurement. The method of measurement was to
optimum H/hc ratio of 0.3. These data are evidently taken from record the proportion of immersion for a given d and plot these
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

Tozzi (1992), and it should be noted that at these flows equi- variables at each section. The (cumulative distribution) curves
librium has not been reached on the 33 m high wall, as shown are nearly symmetrical, giving d50 as the most precise measure
in Fig. 13 in the paper. We suggest that an ‘‘optimal’’ ratio is of depth.
not meaningful under these circumstances and that designers Boes confirms the conclusion that y2 (or clear water depth)
choose a step-height to suit construction practice! It is reveal- is independent of H and, like Matos, notes that the clear water
ing that Fig. 17 suggests that H = 0.5 is the smallest step- depth drops steadily down the spillway. However, Boes’ Fig.
height where the y2/hc ratio remains constant. Note that (Em ⫺ 15 suggests that equilibrium may not have been reached.
Et)/Et does not equal the EDR shown in Fig. 13. If compari- Ohtsu et al. draw attention to the care that needs to be taken
sons between residual energies on stepped and smooth spill- with the positioning of the jump in the stilling basin, and sug-
ways (as used in the EDR) are replaced by comparisons be- gest in their Fig. 24 that the writers’ jump drowned the toe of
tween total specific energy at the crest and residual specific the spillway. There must be another explanation for the dif-
energy at the toe of a stepped spillway, then perhaps a more ference between the two lines in Fig. 24. Ohtsu et al. make
meaningful ratio would be (Em ⫺ Et)/Em, as it would be more no mention of their model scale and their slope is flatter than
likely to detect the convergence to equilibrium. It is noted in ours.
passing that it would be helpful in comparative discussions if On the matter of jump location, Fig. 2 in the paper is the
some standardization of terminology were agreed upon to pre- surface profile of a typical jump drawn to (a distorted vertical)
vent misunderstandings in this very complicated (because of scale; great care was taken to ensure that only horizontal flow
the many variables) field of study. entered the toe of the jump in all cases. Comparison of Fig. 2
Matos’ discussion on air concentrations is interesting but we with Fig. 20 of Ohtsu et al. shows that the jumps were located
cannot comment, as no work on that topic was undertaken, according to their preference. Returning to their text, we can-
except to note the observation that on a 50-m-high spillway not concur with the assumption they make in their (12) that
Matos suggests that equilibrium may not go beyond hc = 3 m. there is a linear relationship between y2 and hc, attractive as
The writers note with pleasure that Matos has also found wavy that might be. The evidence given in Figs. 3, 5, and explicitly
profiles on the spillway (the difficult way) by measuring air 7 in the paper show that y2 = ah bc , where b ⬇ 0.9. This ob-
concentrations and velocities down the slope. Of particular in- servation might have an effect on the conclusion drawn by
terest is the conclusion in Fig. 19 that d, the clear water depths Ohtsu et al. from their Fig. 24.
(and therefore the mean water velocity), become constant in- Tatewar et al. confuse the relationships derived for stepped
dependent of the height of the mean surface. This conjecture spillways of different heights and therefore miss the point of
was offered by the writers in the last paragraph of the discus- Fig. 13 in the paper, which compares EDR with hc for different
sion on Fig. 9 in the paper, where it was stated that ‘‘the mean spillway heights Ymax, and not step-height as they suggest. The
velocity . . . probably reduces as the median depth increases evidence in Figs. 3 and 5 shows that y2 (hence EDR) is in-
(and vice versa) in accordance with intuition.’’ dependent of step-height for the range of flows and step-
Matos confirms the writers’ tentative conclusion on scale by heights considered. What may have misled them is that values
drawing on recently published results. Regrettably, Matos’ for H = 2 m were used in Fig. 13 because that was the largest
equation (3) is in error. Working from the original data-sets, range of flows for which y2 versus hc was recorded (see Fig.
they should read 5); it should be reiterated that over this range of flows the
F2 = 0.196hc0.166 (1:20 scale) (3) sequent depth is independent of H in the range 0.5–2 m, so
0.170
we could have labeled the curve otherwise.
F2 = 0.218hc (1:10 scale) (4) Their Fig. 25 shows a comparison between E1 and y2 values
However the conclusions he draws after presenting these equa- obtained from Fig. 12 for a 58-m-high spillway with (among
tions are in accordance with the data presented in the paper. other information) Tozzi’s data for a 33-m-high spillway. As
Boes’ discussion is based on recent tests on stepped spill- shown in Fig. 13 by comparing results in Figs. 11 and 12,
ways that he conducts with flows up to an impressive 50 m2/ these data for different spillway heights are not commensu-
s, taking the writers’ observations to flow rates at double the rable. If Tatewar et al. plot the power law equation for the
range. Again, as with Matos, corroboration between the studies stepped spillway fitted in Fig. 11, they will observe that it lies
appear which is encouraging. We were aware of the resistivity very close to Tozzi’s data, as is shown by the correct com-
measuring devices mentioned by Boes but decided to stay parisons that appear in Fig. 10 of the paper.
away from what was judged to be a more difficult way of It may be noted in passing from their Fig. 25 that the en-
obtaining equivalent results. Our purpose was not to study air semble of Tozzi’s data group quite well at low y2 and that the
concentrations but to give guidelines to dam-designers in the points for steps smaller than 0.5 m start to diverge from the
sizing of stilling basins at the base of RCC dams. The choice larger for higher flows. This is the reason for not including
of the contact-probe was made because it was seen to yield Tozzi’s data for smaller step-heights in the paper, for as Tate-
more ‘‘robust’’ data. Therefore, in the comparison of specific war et al. observe, for smaller H/hc ratios, the flow experiences
energies, more interest is focused on the residual horizontal smaller frictional resistance. What the findings of the paper
JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000 / 953

J. Hydraul. Eng. 2000.126:947-954.


show in part is that, counter to intuition, there is a remarkably Temple, this correlation invalidates the MEI approach pro-
wide range of geometries where resistance is in fact indepen- posed by the discusser and various associates. However, the
dent of step height. correlation found by Temple was seen by the discusser as an
opportunity to base an estimate of MEI on very simple field
measurements, namely the length of the vegetation h in it’s
upright state or the board drop test introduced by Eastgate
(1966). The discusser extended Temple’s analysis of green
Bermuda grass to all laboratory results published by Ree and
Palmer (1949), Cox and Palmer (1948), and Eastgate (1966).
EFFECT OF RIPARIAN VEGETATION This resulted in a paper that provides two methods to estimate
MEI from simple field methods (Kouwen 1988). These two
ON FLOW RESISTANCE AND
methods gave nearly the same value for MEI for a given test
FLOOD POTENTIALa location.
The statement in the author’s paper that Temple (1987) un-
dertook laboratory experiments in which MEI was correlated
Downloaded from ascelibrary.org by New York University on 05/15/15. Copyright ASCE. For personal use only; all rights reserved.

with vegetation height for a range of growing and dormant or


Discussion by N. Kouwen,2 Member, ASCE dead grass is thus incorrect. Temple (1987) analyzed only a
small part of the discusser’s data. The laboratory experiments
were carried out by Ree and Palmer (1949) and by Cox and
The author has made a significant contribution to the un- Palmer (1948). Furthermore, it was the discusser who back-
derstanding of the effects of vegetation on flow resistance on calculated the values of MEI on the basis of flow measure-
channel banks and floodplains, and is to be congratulated for ments (Kouwen and Li 1980), and not Temple as stated in the
the effort. The discusser is especially pleased to see his own paper.
work incorporated in the analysis. However, there are two
slight misconceptions regarding the origin of (13a,b) and the APPENDIX. REFERENCES
original evaluation of vegetative stiffness, which were erro- Cox, M. B., and Palmer, V. J. (1948). ‘‘Results of tests on vegetated
neously attributed to Temple (1987). waterways and method of field estimation.’’ Miscellaneous Publ. No.
Following a discussion of Temple (1986) by the discusser MP-12, Oklahoma Experiment Station, Stillwater, Okla., 1–43.
(Kouwen 1987), Temple (1987), in his closure, criticized the Kouwen, N. (1987). ‘‘Discussion of ‘Velocity distribution coefficients for
grass-lined channels, by Darrell M. Temple.’ ’’ J. Hydr. Engrg., ASCE,
use of the relationship between k and MEI in (12) because of 113(9), 1221–1224.
the correlation he found between h and MEI. According to Ree, W. O., and Palmer, V. J. (1949). ‘‘Flow of water in channels pro-
tected by vegetated linings.’’ Tech. Bull. No. 967, Soil Conservation
a
May 1999, Vol. 125, No. 5, by Stephen E. Darby (Paper 16426). Serv., U.S. Department of Agriculture, Washington, D.C., 1–115.
2
Prof., Dept. of Civ. Engrg., Univ. of Waterloo, Waterloo, ON, Canada Temple, D. M. (1986). ‘‘Velocity distribution coefficients for grass-lined
N2L 3G1. channels.’’ J. Hydr. Engrg., ASCE, 3(112), 193–205.

954 / JOURNAL OF HYDRAULIC ENGINEERING / DECEMBER 2000

J. Hydraul. Eng. 2000.126:947-954.

You might also like