Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

436 16 Special Manholes

Shockwave Treatment
Each flow perturbation of any basic parameter, including:

• Bottom geometry,
• Cross-sectional geometry,
• Boundary roughness, and
• Discharge variation

has a direct reaction on the generation of a shockwave. Besides the definition of


the shockwave geometry, the shockwave reduction deserves attention. A thorough
design on any hydraulic structure thus includes a downstream flow without flow
concentrations and with a structural economy. If these items are not attained, an
intermediate stilling basin would be appropriate to reduce the disadvantages of high-
speed flows. From the hydraulic point of view, such a design would be forced or even
poor, and solutions to be suggested below aim at an entirely supercritical flow across
any hydraulic element. Such solutions are advantageous from both the hydraulic and
the structural views.
Currently, a number of methods for shockwave reduction are available, but these
are based on assumptions such as:

• Neglect of boundary layer effects,


• Hydrostatic pressure distribution,
• Omission of bottom and friction slopes, and
• Investigation for design flow only.

Such simplifications are ruled out provided the design is based on either:

• Serious two-dimensional numerical modelling of supercritical flow, or


• Flow analysis by sufficiently large physical models to inhibit scale effects.

Currently, scale models are an economic and reliable approach to solve these
problems, and numerical models have received less attention.
In the following, supercritical flows of typical structures relating to sewer
hydraulics are discussed, such as constrictions, expansions, bends and junctions.
First, the flow patterns of the untreated channel are presented, and structural means
to reduce shockwaves are then outlined. The indications refer normally to rectan-
gular channels, because U-shaped and circular channels have not yet received a
systematic attention. Some results pertinent to sewer structures follow in Sect. 16.4.
Few results were available until the 1990s; this lack was particularly due to the mod-
eling of supercritical channel flow. Whereas most of the studies were based on a
poor approach flow whose surface was perturbed by significant shockwaves prior to
enter the test reach, the present Author developed with his colleagues the so-called
jet-box allowing for an almost perfect approach flow absolutely free of shockwaves
and velocity concentrations.
16.3 Supercritical Flow 437

16.3.3 Channel Contraction


Conventional Design
A channel contraction (German: Kanalverengung; French: Contraction de canal)
can be geometrically designed either funnel-, fan- or nozzle-shaped. The follow-
ing is restricted to polygonal-funnels, because curved contraction walls are not a
standard in sewer hydraulics (Fig. 16.22).
A hydraulically sound designed contraction is characterized by a monotonically
rising surface in the constricted portion, and a downstream channel practically free
of surface waves. In the approach flow region ➀, the flow depth is h1 and the veloc-
ity V1 , such that the approach flow Froude number is F1 = V1 /(gh1 )1/2 . In analogy
to the abrupt wall deflection (Sect. 16.3.2), the origin of perturbation is the contrac-
tion angle θ at point A. In region ➁, the characteristic parameters are thus h2 and
according to Eq. (16.55) V2 = V1 . Point D generates negative shockwaves due to the
negative wall deflection, and the downstream channel has the typical configuration
with crosswaves. The flow uniformity is thus poor in the untreated channel.
Ippen and Dawson (1951) suggested a design in which the contraction angle θ
is selected such that the positive shock front is directed to the origin of the negative
shock front, and points C and D in Fig. 16.22a coincide (Fig. 16.22b). According to
the wave interference principle, the downstream waves should then be eliminated.
A contraction structure so designed has three regions, namely the approach flow
region ➀, the contraction region ➁, and the downstream region ➂, with the corre-
sponding flow depths h1 , h2 and h3 . Applying successively the relations derived in
Sect. 16.3.2, the design contraction angle θ ( < 10◦ ) can be given as (Hager 1992)

Fig. 16.22 (a) Schematic


flow in polygonal contraction
and (b) design in plan (top)
and section (bottom) with (–)
wall and (- -) axial profiles
(Ippen and Dawson 1951)
438 16 Special Manholes
 
b1 1
arctan θ = −1 . (16.62)
b3 2F1

To inhibit choking (German: Strömungszusammenbruch; French: Effondrement


de l’écoulement) of downstream flow, the condition F1 > 2 must be satisfied.
Equation (16.62) can be used to demonstrate a serious disadvantage of all flows
designed with the interference principle. A minor deviation from the design con-
ditions leads to serious shockwaves, i.e. the design flexibility is extremely narrow.
In addition, Reinauer and Hager (1998) have demonstrated experimentally that the
wave interference principle as applied in optics cannot be transmitted to hydraulics,
mainly because of the nonlinear wave characteristics. Whereas waves in optics have
practically no transverse extension, the wave lengths in hydraulics are finite and
have the order of a flow depth.

Novel Design
Reinauer and Hager (1998) investigated experimentally supercritical flow in a chan-
nel contraction, for bottom slopes up to 30◦ . Figure 16.23 shows a definition sketch
in the horizontal channel without any bottom elements. These have been pro-
posed for chute contractions with a large discharge and are not relevant for sewer
contractions.
At the contraction point A a shockwave of shock angle β 1 is generated due to the
abrupt wall deflection of angle θ . The shock is propagated to point B in the channel
axis and reflected to the wall again at point C. At the contraction end point E, a nega-
tive wave forms, resulting in a complicated wave pattern in the downstream channel.
In the side view the wall (subscript w) and the axial (subscript a) surface profiles are
relevant. Three shockwaves can be distinguished: Wave 1 of height h1 and wave 3
of height h3 along the wall, and the axial wave 2 of height h2 . Downstream of wave
1, the flow depth hp in the wall region is nearly constant up to the height he at the

Fig. 16.23 Horizontal


channel contraction (a) plan
(b) section with surface
profiles along (–) wall and
(- -) channel axis
16.3 Supercritical Flow 439

contraction end. The velocity across the contraction is essentially constant, as for a
flow behind an abrupt wall deflection (Sect. 16.3.2).
The shock waves in a channel contraction depend mainly on the:

• Approach flow Froude number Fo = Vo /(gho )1/2 with Vo = Q/(bo ho ),


• Width ratio ω = be /bo ,
• Wall deflection angle θ , and
• Bottom slope So .

Effects of approach flow depth, i.e. scale effects are inhibited provided the mini-
mum approach flow depth is ho = 50 mm. For waves 1 and 2, the width ratio is
insignificant, such that the approach flow shock number So = θ Fo is the govern-
ing parameter. Shockwaves for So < 0.5 are weak, and they become overforced for
So > 2.
A shockwave may be described with the coordinates (x;h) of location and wave
maximum. With the bottom angle α in [deg] and the contraction angle θ in [rad] the
relative heights of waves 1 − 3 may be expressed as

 2
1
Y1 = h1 /ho = 1 + √ So , (16.63)
2

Y2 = h2 /ho = (1 + 2So )2 , (16.64)

Y3 = h3 /ho = ω−1 + 1.8S1/2


o − 0.2 α .
0.6
(16.65)

Therefore, the contraction angle and the bottom slope have an effect only on wave 3.
The locations of wave maxima are more complex because they involve a drawdown
of the free surface profile along the sloping channel. The application limits of Eqs.
(16.63), (16.64), and (16.65) are 0.2 < ω < 1, α < 45◦ , and So < 2 (Reinauer and
Hager 1998).
Equation (16.63) gives a height of the shock wave identical with that of an abrupt
wall deflection. For the maximum value So = 2 recommended, Y1 ∼ = 6 resulting in a
significant wave height. Wave 2 with Y2 ∼ = 15 is much higher, however. For a typical
parameter combination ω = 0.5, and α = 10◦ , the result for wave 3 is Y3 = 3.75.
The important conclusion from this study is that the effect of bottom slope decreases
the wave height, and a conservative approach assuming α = 0 gives results on the
safe side. Typically, a shock number of So = 1 should be selected for design, and
the optimum contraction rate is between 0.5 < ω < 0.80. For ω < 0.5, the flow tends
to be overforced, and contractions with ω > 0.8 are often not justified because of
additional cost. The design parameter for the maximum wave height is the maximum
discharge QM . All other discharges Q < QM result in smaller waves.
Choking flow in contractions is a significant design limitation. If the approach
flow Froude number Fo or the contraction rate ω are too small, then the supercritical
flow structure breaks down, associated with a hydraulic jump in the contraction of
440 16 Special Manholes

Fig. 16.24 Choking flow in a channel contraction for Fo = 2.40

which the toe is located in the approach flow channel. Figure 16.24 shows a typical
sequence between the limiting downstream flow and a stable hydraulic jump in the
contraction.
Starting from a supercritical flow, Fo can be reduced up to the occurrence of
incipient choking, with Fo = Fo– . Then, the flow depth at point E is equal to the
critical depth, and the hydraulic jump establishes, with a transition from super-
critical approach flow to subcritical contraction and supercritical downstream flow.
A choked contraction has a much larger flow depth than the maximum wave height.
Therefore, choking of any hydraulic structure that was designed for supercritical
flow is dangerous, because the flow structure assumed is significantly different from
the choked flow pattern. Once a structure has a choked flow, a Froude number
Fo+ much larger than Fo– is needed to blow out the hydraulic jump. For a given
approach flow Froude number in an almost horizontal channel, the limit (subscript
L) contraction ratio is
 3/2
3
ωz = Fo . (16.66)
2 + F2o

A contraction flow chokes if ω < ωz . For bottom slopes α > 5◦ choking for ω > 0.5
is unlikely to occur (Reinauer and Hager 1998).

16.3.4 Channel Expansion


A channel expansion (German: Kanalerweiterung; French: Expansion de canal) can
be geometrically arranged in various shapes. Mazumder and Hager (1993) stated
that only the abrupt expansion and the so-called reversed expansion are relevant.
Whereas the latter is applied in hydraulic structures for larger discharges, sewer
hydraulics involve normally abrupt expansions (Fig. 16.25). The following refers to
a rectangular channel in which the friction slope is nearly compensated for by the
bottom slope. The approach (subscript o) flow depth is ho and the approach flow
16.3 Supercritical Flow 441

Fig. 16.25 Abrupt channel


expansion. (a) Section with
axial surface profile ha (x) and
wall profile hw (x). (b) Plan
with (· · ·) profiles of shock
fronts

Fig. 16.26 Free surface downstream of abrupt channel expansion with bu /bo = 3, Fo = 2 and
ho = 96 mm. (a) Photo, (b) schematic plot (Hager and Mazumder 1992)

Froude number Fo = Vo /(gho )1/2 . The approach flow width bo increases abruptly to
the downstream (subscript u) width bu at the expansion section.
Following an expansion, the free surface draws down in the approach flow
channel (Fig. 16.25a). At the expansion section, the flow has nearly a rect-
angular cross-section, with vertical sides, such that its pressure distribution is
non-hydrostatic. This jet expands over the downstream width bu and delimits dead-
water zones in either corner. At the impinging points A and B of the forward flow
fronts, shockwaves are generated similar to those due to a wall deflection. Contrary
to the latter, the approach flow to the shockwaves is already perturbed due to flow
expansion, and the relation between the maximum (subscript M) wall wave height
hwM and the approach flow depth ho is complicated. Downstream from the impact
points A and B, the waves are reflected and the various flow zones typical for super-
critical flows are again visible (Fig. 16.26). In analogy to Fig. 16.21, region ACB is
442 16 Special Manholes

influenced by the approach flow only, resulting in the maximum flow depth at points
A and B. In region ACE, the average flow depth is larger than upstream from point
A, with an axial maximum flow depth haM at point C. Downstream, the typical
crosswave pattern appears, with successive wave minima and maxima along both
the channel axis and walls of which all are smaller than upstream.
The maximum flow depths that are determining for freeboard design occur at the
first wave both along the channel axis (subscript a) and the channel walls (subscript
w). An analysis of these maxima is thus more relevant than information on the com-
plete flow surface. Figure 16.27 refers to the axial profile Ya (X) with Ya = ha /ho
and the wall profile Yw (X) with Yw = hw /ho in which the dimensionless stream-
wise coordinate is stretched with Fo to X = x/(bo Fo ) for various expansion ratios
β e = bu /bo ≥ 1. The axial profile is seen to be independent of β e and Fo and may be
expressed up to the crossing point C as

Ya = 0.2 + 0.8 exp (− X 2 ). (16.67)

The wall profile Yw (X) can be represented as

Yw = YwM τ exp (1 − τ ) (16.68)

with the maximum wall flow depth YwM = hwM /ho for 1.8 < β e < 6 as

YwM = 1.27βe−0.4 . (16.69)

The normalized streamwise coordinate

X − Xm
τ= (16.70)
XM − X m

depends on the locations of the wave extrema

Xm = (1/6)(βe − 1), XM = 0.52 βe0.86 . (16.71)

Fig. 16.27 (a) Axial surface profile, and (b) wall surface profile for abrupt channel expansion for
1 < Fo < 10 and 1.8 < β e < 6 (Hager and Mazumder 1992)
16.3 Supercritical Flow 443

The location Xm refers to the transition between the dead-water zone and the wave
beginning with the minimum (subscript m) flow depth hwM = 0, and XM is the point
of maximum wave height (points A and B of Fig. 16.26).

Example 16.5 Given a rectangular channel of approach flow velocity Vo =


6 ms–1 and approach flow depth ho = 0.60 m. Describe the downstream flow
pattern for an abrupt width increase from bo = 0.8 m to bu = 3.0 m.
With Vo = 6 ms–1 and ho = 0.60 m, the approach flow Froude number is
Fo = 6/(9.81 · 0.6)1/2 = 2.50, and the width ratio is β e = 3.0/0.8 = 3.75. The
locations of extreme wave heights are Xm = (1/6)(3.75–1) = 0.46, and XM =
0.52 · 3.750.86 = 1.62, according to Eqs. (16.71). The normalized length coor-
dinate is thus τ = (X – 0.46)/1.16. The maximum wall wave height is YwM =
1.27 · 3.75–0.4 = 0.75 from Eq. (16.69), and the equation of the wall profile is

Yw = 0.75[(X − 0.46)/1.16] exp [1 − (X − 0.46)/1.16] (16.72)

from Eq. (16.68). The results are thus xM = bo Fo XM = 0.8 · 1.62 · 2.5 =
3.2 m, and the maximum wall flow depth is hwM = 0.75 · 0.60 = 0.45 m.

Further indications refer to the flow surface, the shock fronts, the transverse sur-
face profiles and the velocity profiles. The dead-water zone downstream from the
expansion section can be suppressed by linear insets up to point xm according to
Eq. (16.71). Then, the abrupt expansion is replaced by a gradual expansion, without
a potential for sedimentation of the sewage.
The internal flow characteristics of expanding flows up to the point of shock-
wave development has been investigated by Hager and Yasuda (1997). Based on
Eqs. (16.40), (16.41), (16.42), and (16.43) the 2D shallow water equations were
demonstrated to tend asymptotically to the 1D unsteady flow equations, and a flow
analogy between steady two-dimensional and unsteady one-dimensional flows was
established.

16.3.5 Channel Bend


Weak Bend Flow
A channel bend (German: Kanalkurve; French: Canal courbé) of deviation ζ
involves a supercritical flow similar to a wall deflection with a positive shockwave
along the outer and a negative shockwave along the inner wall, respectively. Further
downstream the wave maxima and minima alternate and typical crosswaves are
generated. Knapp (1951) discussed such flow and set up a design procedure.
444 16 Special Manholes

Fig. 16.28 Supercritical flow in a channel bend (a) schematic flow process, (b) angle β s versus
approach flow Froude number Fo and relative bend curvature b/R (Hager 1992)

The flow pattern in a rectangular horizontal channel containing a frictionless fluid


can be simplified as follows. Extreme wave heights occur at locations with angles
θ , 2θ and so on (Fig. 16.28). At the end of the bend, this oscillating surface pattern
does not stop but continues into the tailwater channel.
The elementary shockwave angle is β s = arcsin(Fo– 1 ). Knapp (1951) related this
angle to the shockfront angle θ and to the relative centerline bend curvature b/R as
 
b/R
θ = arctan . (16.73)
(1 + 2b/R) tan βs

Figure 16.28b indicates that θ increases as both Fo and b/R increase. For small
values of Fo , the curves are dashed because of the undular surface pattern. For
larger approach flow Froude numbers resulting in direct shockwaves provided b/R
< 1/2, Eq. (16.73) can be written as

b/R
tan θ = F2o − 1 ∼
= (b/R)Fo . (16.74)
1 + (1/2)(b/R)

The extreme wave heights (subscript e) may be obtained with the energy prin-
ciple, and Figure 16.29 shows the result of Knapp (1951). For (b/2R)Fo2 < 1 the
extreme values ye = he /ho may be approximated as

 2
1 2
ye = 1 ± (b/R)Fo , (16.75)
2
16.3 Supercritical Flow 445

Fig. 16.29 Curve flow in


rectangular channel, extreme
flow depths ye = he /ho versus
Fo = Vo /(gho )1/2 and relative
centerline curvature b/R
(Hager 1992)

with the +sign relating to the wave maximum and the –sign to the wave minimum,
respectively. In analogy to the shock number S = θ F of an abrupt wall deflection,
one may define the bend number as B = (b/R)1/2 F. The effects of the square root of
relative bend radius (b/R)1/2 , and approach flow Froude number are thus contained
in the bend number. Compared to the abrupt wall deflection, the effect of b/R is
smaller. Note that the effect of the approach flow Froude number Fo is significant
and shockwaves are high for medium values of Fo .

Example 16.6 Given a rectangular channel with an approach flow velocity


Vo = 7 ms–1 , bo = 1.2 m and ho = 0.50 m. How high should be the sidewalls
for a deflection angle of ζ = 35◦ , if the centerline radius is R = 8 m?
With Vo = 7 ms–1 and ho = 0.50 m, the approach flow Froude num-
ber is Fo = 7/(9.81 · 0.50)1/2 = 3.16, and the relative centerline curvature
amounts to b/R = 1.2/8 = 0.15 < 0.50. Accordingly, the term (1/2)(b/R)
Fo2 = 0.5 · 0.15 · 3.162 = 0.75 < 1, and ye = (1±0.75)2 , thus yM = 3.06
and ym = 0.06, and the maximum wall flow depth is hM = 3.06 · 0.5 =
1.53 m and the minimum flow depth is hm = 0.06 · 0.5 = 0.03 m. From
Fig. 16.29 one finds for Fo–1 = 3.16–1 = 0.32 values yM = 2.3 and
ym = 0.15, i.e. values closer to 1 than according to the approximations.
Further, from Eq. (16.74), tan θ = 0.15 · 3.16 = 0.47, and θ = 25◦ . The first
wave extrema are thus located within the curved channel portion.

The effects of bottom slope and transverse slope have not yet been systemat-
ically investigated. Results for conduit bends follow below. Figure 16.30 shows
photographs of flows in horseshoe profiles. The transverse free surface slope can
be considerable, and choking of such tunnels for large discharges is possible. Then
a complex air-water flow may set up, with a hydraulic jump in the upstream portion,
and pressurized flow in the tailwater portion. The limit conditions for choking flow
in curved channels should be analyzed, therefore.
446 16 Special Manholes

Fig. 16.30 Flow in closed


conduit of large bottom slope
(a) upstream view,
(b) downstream view

Strong Bend Flow


Reinauer and Hager (1997) expanded the approach of Knapp (1951). Figure 16.31
shows flows in bends of rectangular channels in which large wall shockwaves may
be observed along the outer channel wall, whereas the flow depth along the inner
channel wall may reduce to zero due to the centrifugal force.
The velocity fields of bend flow reveal interesting flow features. From detailed
observations, one may conclude that the (Reinauer and Hager 1997):

• Tangential velocity component remains almost constant along the curve, with a
velocity identical to the approach flow velocity for Fo > 3.
• Velocity distribution in the vertical direction is also practically constant, except
for the boundary layers along the bottom and the walls.

Figure 16.32 shows a definition plot for bend flow in a rectangular channel,
with the wave maxima (subscript M) and wave minima (subscript m) at location
θ , 2θ , ... The approach flow velocity is Vo , the approach flow depth ho , such that
Fo = Vo /(gho )1/2 is the approach flow Froude number. With Ra as the centerline
radius of curvature, or b/Ra as the relative bend curvature, the significant parameter
for supercritical bend flow is the bend number B = (b/Ra )1/2 F, with F = Fo . Based
on the existing observations and own experiments, the maximum wave height YM =
hM /ho depends exclusively on the bend number as (Reinauer and Hager 1997)

YM = (1 + 0.4B2 )2 , B ≤ 1.5, (16.76)


16.3 Supercritical Flow 447

Fig. 16.31 Surface of supercritical bend flow for Fo = (a) 4, (b) 6. (· · ·) Shock front

Fig. 16.32 Definition plot


for supercritical bend flow in
plan view
448 16 Special Manholes

YM = (1 + 0.6B)2 , B > 1.5. (16.77)

For relatively small B, Eq. (16.76) is an approximation of Knapp’s approach and


the present predictions are almost identical. For large values of B, the effect of the
bend number is linear, and Eq. (16.77) has to be applied. The change of Eq. (16.76),
and (16.77) could not be attributed to a physical feature, such as flow separation
from the inner bend wall, or wave breaking.
The location tan θ M of the wave maximum is governed by the parameter (b/Ra )F
instead of (b/Ra )1/2 F, in agreement with Eq. (16.74). The experimental results are

tan θM = (b/Ra )F, for (b/Ra )F ≤ 0.35, (16.78)

tan θM = 0.6[(b/Ra )F]1/2 , for (b/Ra )F > 0.35. (16.79)

Accordingly, Ippen and Knapp’s approach applies only to flows of relatively small
bend number.
The location and height of the wave minimum (subscript m) can also be
predicted as

Ym = (1 − 0.5B2 )2 , (16.80)

tan θm = 2(b/Ra )F. (16.81)

For Ym < 0, the flow is completely separated from the wall; Reinauer and Hager
(1997) predicted the geometry of the separated flow region. Figure 16.33 shows
plan views of bend flows and Fig. 16.34 refers to upstream views. In the latter case
the bend flow is seen to occupy only a portion of the channel width, and the design
may be highly uneconomic. Choking was not a problem in the rectangular channel,
but it may be dangerous in pipes. The main results are presented below.
The wave profiles along the inner and outer bend walls have been further ana-
lyzed. Figure 16.35 shows the normalized profile τ M (θ /θ M ) along the outer wall,

Fig. 16.33 Plan view for


bend flow with b/Ra = 0.07
and Fo = (a) 4, (b) 8
16.3 Supercritical Flow 449

Fig. 16.34 Upstream views of bend flow for b/Ra = 0.144 and Fo = (a) 4, (b) 8

Fig. 16.35 Generalized wall flow profiles along (a) outer and (b) inner bend walls for b/Ra = 0.07
and Fo = (×) 2.5, () 3, () 4, () 6, (•) 8

with τ M = (h – ho )/(hM – ho ), and similarity is noted up to θ /θ M = 1.25. For larger


deflection the experimental curves spread, depending on Fo . For θ up to about 50◦
the outer wave profile follows

τM = [ sin (θ/θM )]1.5 . (16.82)

The profile along the inner bend wall τ m (θ /θ m ) corresponds to the profile along
the outer wall. With τ m = (h – ho )/(hm – ho ), a single curve may be plotted up to
θ /θ m = 1.2. For smaller values of Fo , the wave trough is followed by a wave crest,
whereas flow separation from the inner wall results for larger Fo . Equation (16.82)
is verified for relative curvatures b/Ra = 0.144 and 0.31 in Fig. 16.36a.
The effect of partial bend flow with a deflection angle ζ smaller than θ M is
shown in Fig. 16.36b). At the bend end, the rise of the surface profile according to
Eq. (16.82) stops almost abruptly. Then, the maximum flow depth is predicted from
Eq. (16.82) by setting θ = ζ .
450 16 Special Manholes

Fig. 16.36 (a) Outer wall profile τ M (θ/θ M ) for b/Ra = 0.144 and 0.31, (–) Eq. (16.82), (b) partial
bend flow Y(θ) for θ = 30◦ (light) and θ = 50◦ (solid). Notation Fig. 16.35

Fig. 16.37 Choked bend flow with forward flow along the outer wall and recirculated flow along
the inner bend wall. (a) Upstream view, (b) and (c) plan views

Figure 16.37 shows submerged bend flow with a supercritical approach flow.
The forward flow concentrates along the outer wall, and a large recirculation zone
is observed along the inner bend wall. Such flows are highly unstable and have to
be avoided because of the poor hydraulic performance. Currently, such flows have
not yet received systematic treatment, however.

Tunnel Bend Flow


Bend flows may choke in a tunnel if its capacity is too small to discharge free surface
flow. These flows were analyzed experimentally by Gisonni and Hager (1999) rela-
tive to bottom outlets. Whereas a bottom outlet is normally deflected by an angle of
45◦ , a sewer usually has a 90◦ bend. In the following, only results for a 45◦ deflec-
tion angle are presented, which apply approximately also to the 90◦ -deflected sewer
bend.
16.3 Supercritical Flow 451

Fig. 16.38 Flow downstream


of 45◦ -deflected tunnel bend
for yo = 0.24 and Fo = (a)
5.2, (b) 5.8, (c) 11.8 with
general view (top), plan view
on tailwater pipe (center or
bottom), and side view
(bottom of c)

Figure 16.38 relates to the straight tailwater reach downstream of the bend, for
an approach flow filling ratio of yo = ho /D = 0.24 and a range of approach flow
Froude numbers Fo = Q/(gD4 ho )1/2 (Chap. 6), where D is the tunnel diameter and
ho the approach flow depth. The relative tunnel radius was R/D = 3. For Fo = 5.2
(Fig. 16.38a), so called stratified tunnel flow was observed, for which the air phase
is above the water flow. As Fo = 5.8 is reached (Fig. 16.38b), the transition from
stratified to annular flow occurs just downstream of the tunnel bend. Upon further
increasing to Fo = 11.8 (Fig. 16.38c), the entire tailwater reach was in the annular
flow regime (Chap. 5).
Figure 16.39 relates to a higher filling ratio of yo = 0.61, for which the flow struc-
ture is similar as for yo = 0.24 except that the transition Froude number between
stratified and annular flows is reduced. Figure 16.39a shows the transverse sur-
face profile consisting mainly of black water along the outer, and a white water
front along the inner tailwater sides. If either yo or Fo are increased, then the pipe
chokes resulting in pressurized two-phase flow. The transition occurs for Fo ∼ = 2.75
(Fig. 16.39b), and (Fig. 16.39c) shows annular flow for Fo = 4.2 including a long
vortex spanning from the bottom of the bend outlet toward the top of the pipe end.
The following relates to the main hydraulic features of these flows.
452 16 Special Manholes

Fig. 16.39 Tunnel flow downstream of 45◦ -deflected bend for yo = 0.61 and Fo = (a) 2.0, (b)
2.75, (c) 4.2

Figure 16.40 is a definition plot for the flow considered, including the approach
flow water discharge Q and the approach flow air discharge Qa , the distance d1 from
the bend end to the first wave maximum hM and the height h2 of the second wave
maximum. As demonstrated in Chap. 7, the sequent depth ratio of flow in a straight

Fig. 16.40 Definition plot


for supercritical tunnel bend
flow with (a) plan, (b)
streamwise section
16.3 Supercritical Flow 453

circular pipe is Y = F10.90 , with subscript 1 relating to the supercritical approach


flow. Choking from free surface to pressurized pipe flow occurs nominally for
h2 = D. Inserting this condition in the previous sequent depth relation leads to the
so called choking number
1/2
C = F1 /(D/h1 ) = Q/ gD3 h21 . (16.83)

Note that all three governing parameters discharge Q, pipe diameter D and approach
flow depth h1 have an almost linear effect on C. The choking number so defined
includes both the filling ratio of the approach flow and its dynamics as expressed by
the approach flow Froude number.
A tunnel bend generates shockwaves for supercritical flow whose effect reduces
the capacity as compared to a straight tunnel reach. The choking number of a tunnel
bend is defined from Eq. (16.83) as Co = Fo /(D/ho ) and may be considered the rele-
vant parameter for these flows. The relative wave maximum YM = (hM /ho )(ho /D)2/3
in the tailwater reach (Fig. 16.40b) may be expressed as (Gisonni and Hager 1999)

YM = 1.10 tanh (1.4Co ), Co > 0.80. (16.84)

For Co < 0.80, the tunnel flow was subcritical because a hydraulic jump formed at
the bend inlet. For Co > 1, the maximum relative wave height is almost constant at
YM = 1.1 indicating hM = 1.1(ho D2 )1/3 , corresponding to a maximum filling ratio of
hM /D = 1.1(ho /D)1/3 . The distance d1 from the bend end to wave 1 may be expressed
as d1 /D = 4.8(Co – 0.8)1/2 . For incipient choking, the first wave was located some
2 pipe diameters downstream from the tunnel end. Wave 2 downstream of the max-
imum wave is always less high, i.e. h2 < hM , and further waves were hardly visible.
No effect of a small streamwise bottom slope on this and the following relations was
noted.
If annular flow is established downstream of the bend tunnel, there is a transition
to stratified flow at a length da once the rotational component is insufficient. The
relative distance was found to be

da /D = 0.85(Co − 1)2 , Co > 1. (16.85)

Accordingly, the tunnel just issues a stratified flow for Co = 1, such that this is the
condition for stratified tunnel bend flow.
The distance of the first wave crest from the tunnel outlet was further measured,
resulting in

d1 /D = 4.8(Co − 0.8)1/2 , Co > 0.8. (16.86)

For incipient choking, the first wave was located about 2 pipe diameters downstream
of the bend end, with a corresponding filling ratio of hM /D = 0.97(ho /D)1/3 .
The air discharge Qa across the tunnel bend was also measured. This discharge
was observed to increase with Co and to decrease with the filling ratio, because the
454 16 Special Manholes

cross-sectional area for air flow reduces. Accordingly, the usual air-water discharge
ratio as presented in Chap. 5 was generalized to B = (Qa /Q)/(1 – yo )2 . The tests
indicated that B increases essentially with Co if Co > 1. For lower values of Co there
is also air flow but mainly due to free surface entrainment, as in open channel flow.
The data were fitted to

B = 3 tanh [3(Co − 1)], Co > 1. (16.87)

indicating a maximum of Qa /Q = 3(1 – yo )2 for Co > 1, providing typically values of


Qa /Q = 1 – 2. This is a comparatively large air discharge ratio asking for a separate
air supply across the upstream manholes. It was also observed that cover plates as
introduced in Sect. 16.3.7 for bend manholes do not at all improve the flow pattern
across a bend tunnel.

16.3.6 Channel Junction


Flow Description
Figure 16.41 refers to a channel junction (German: Kanalvereinigung; French:
Jonction de canal) with a straight through (subscript o) and a lateral (subscript z)
branch under the junction angle δ. Further, the widths of the upstream and down-
stream branches are equal to bo = bu , whereas bz ≤ bo . Because the flow close to
the junction point P is only considered, the effects of bottom slope and wall friction
are neglected. The general case of a junction structure with three different widths
bo , bz , and bu , and two different branch angles δ o and δ z has not yet been analyzed.
Regarding the branch discharges, three cases can be discussed:

• Discharge Qo > 0 and no lateral discharge,


• Discharge Qz > 0 and no upstream discharge, and
• Discharges Qo > 0 in the upstream, and Qz > 0 in the lateral branch.

Fig. 16.41 Channel junction with (a) upstream discharge only, (b) lateral discharge only
16.3 Supercritical Flow 455

For junctions with Qz = 0, the governing parameters are (ho ,Fo ). Downstream from
the junction point P, the flow may laterally expand and impinge on the lateral branch
with a stagnation point I (Fig. 16.41a). Depending on the junction angle δ, a small
flow portion is directed into the lateral branch but the main discharge portion con-
tinues in the downstream branch. By flow impact on the wall between points I and
W, a standing wave is formed, referred to as wall wave A with a maximum flow
depth hMA at location xMA . The end (subscript e) of the wave is at location xeA .
For junctions with Qo = 0, i.e. lateral discharge only, the flow expands from
the junction point P and impinges on the opposite wall, with a stagnation point K.
Depending on the junction angle δ, a smaller or larger discharge portion flows into
the upstream branch and forms a recirculation zone. Downstream from point K, a
standing wave is formed, referred to as wall wave B. Its maximum flow depth hMB
is at location xMB , and its end is at xeB . Opposite from wave B, and starting at point
W a recirculation zone is established with a stagnation point U in the downstream
branch.
Both waves A and B can be either compact or wall-type waves (Fig. 16.42).
A compact wave is continuous and smooth with a nearly hydrostatic pressure
distribution. These waves are typical for weak junction flow. A wall-type wave
rises up the wall, is narrow compared to the approach flow width and results from
overforced flow. Waves A and B may be either pure water waves or highly aerated
two-phase flows with a considerable spray development, depending on the Froude
number and the absolute velocity. Supercritical junction flows can thus be highly
spatial and then tend to flow concentrations in the downstream branch, combined
with large separation regions.
A channel junction with discharges in both the upstream and lateral branches
can be defined with the parameters (ho ,Fo ) and (hz ,Fz ), respectively (Fig. 16.43).
Normally, wave A is suppressed by the lateral discharge but wave C is formed,
starting from the junction point. Wave C is directly related to the merging of
two supercritical flows of different direction and different flow depth. It can be
interpreted as a vertical movement of two horizontal flows representing the only
possibility of momentum exchange between two high-velocity flows. Wave C has a
maximum depth hMC at location xMC , and the shock front has an angle θ relative to

Fig. 16.42 Wall wave types (a) compact and (b) wall-type waves
456 16 Special Manholes

Fig. 16.43 Simple channel junction with supercritical approach flows, (a) plan, (b) section with
wave structure and (- - -) recirculation zone

the downstream direction. Note the analogy between the abrupt wall deflection and
the junction structure regarding wave C.
The shock front of wave C impinges the opposite wall at point K, generating
wave B downstream from it, with hMB as maximum wave height at location xMB .
On the opposite wall, the flow separates (subscript s) between point W and the
stagnation point U, with the maximum width ys at location xs . Wave B is reflected
into the downstream branch, to generate wave D. Depending on the ratio of wave
amplitudes hMD /hMB , the undular flow character is rapidly or slowly dampened in
the downstream branch.
Based on the previous description of general junction flow, a complex flow
pattern is always established. It is influenced by a large number of parameters
such as the ratio of branch flow depths ho /hz , the branch Froude numbers Fo and
Fz , the width ratio bo /bz and the junction angle δ. In addition, various junction
geometries could be considered, with rounded corners, bottom drops and changes in
the boundary roughness. In the following, some results originating from systematic
experimentation relating to the simple junction as shown in Fig. 16.43 are presented.

Limit Conditions
Incipient choking (German: Grenzzustand; French: Commencement d’effondre-
ment) corresponds to the transition between super- and subcritical junction flows.
Depending on whether incipient choking appears in one or even in two branches of
a junction, supercritical flow is maintained in one or even no branch. Knowledge of
the limit (subscript L) conditions is essential because choked junction flow behaves
completely different from the supercritical junction flow described previously.
Figure 16.44 shows the stages of flow choking in one branch alone. Both con-
ditions are generated provided the flow at the stagnation point K for the through
branch, and at the stagnation point I recirculate. The limit condition depends thus
16.3 Supercritical Flow 457

Fig. 16.44 Limit conditions in channel junction for (a) through and (b) lateral branch, with phases
from incipient choking to stable hydraulic jump

on the exact flow structure and the junction geometry. Figure 16.44 relates to the
four main phases of choking, from incipient choking to the stable hydraulic jump
in one branch, and supercritical flow in the other branch. If the Froude number is
further reduced, choking appears also in the latter branch, resulting in fully choked
junction flow.
The experimental data regarding the limit (subscript L) condition of incipient
choking may be expressed as (Schwalt 1993)

FLo = 2 + (δ/120◦ )Fz Y −1 , (16.88)

FLz = 2 + (1/8)(Fo + 5)Y. (16.89)

Both limit Froude numbers thus depend linearly on the Froude number of the neigh-
bour branch and on the flow depth ratio Y = ho /hz . FLo varies with the junction angle
δ in addition, whereas FLz does not. An absolute lower limit for both FLo and FLz
is F = 2, because supercritical approach flow for F < 2 is always choked. These
flows typically generate undular hydraulic jumps (Chap. 7). Prior to the design of a
junction structure, one has thus to verify both conditions Fo > FLo and Fz > FLz .

Wave Geometry
The main waves of a junction subjected to supercritical flow are waves B, C and
D (Fig. 16.43). Usually, wave A cannot be observed if both branches operate, and
wave D and all successive downstream waves are smaller than the determining
wave B. Particular attention receive thus waves B and C.
Each wave can be defined by its beginning (subscript a), maximum (subscript M),
and end (subscript e). In addition, the width yMB of a wall wave at the maxi-
mum location can be defined. Further, the main characteristics of the separation
(subscript s) zone are of interest. With x as the streamwise coordinate, y as the
transverse coordinate measured from the wall opposite of the lateral branch, and
h as the flow depth, these parameters generally vary with the basic parameters
Fo = Vo /(gho )1/2 , Fz = Vz /(ghz )1/2 , Y = ho /hz , and β = bz /bo . Based on systematic
experimentation, Schwalt (1993) found for
458 16 Special Manholes

Wave C
xMC − xP
tan θ = = (1/2)δ(Fz /Fo )2/3 Y −1 , (16.90)
bo − yMc
xMC − xP
LMC = = 4.3(f − 2), (16.91)
h̄ cos θ
hMC
ZMC = − 1 = 1 + 0.25(f sin δ)2 . (16.92)

Here, xP is the location of junction point P, θ the shock angle, f = 2Fo Fz /(Fo + Fz )
the determining Froude number, h̄ = (ho hz )1/2 the determining approach flow depth,
and b̄ = (bo bz )1/2 the determining approach flow width. With Δ = 0.55 + 0.05δ ◦ ,
Schwalt (1993) found for the characteristics of

Wave B
xaB − xP 
LaB = = (1/Δ) Fo /F1/3 z Y 2/3 , (16.93)
bo
xMB − xP
LMB = = 1.65LaB , (16.94)
bo
xeB − xP
LeB = = 1.35 cos δ Fo (Fz Y)1/3 + 5 , (16.95)
(bh)1/2
yMB
BMB = = 0.3(0.1 + sin δ)Fo F1/3
z , (16.96)

hMB
ZMB = − 1 = 0.25f 2 . (16.97)

Wave D starts at location (Schwalt and Hager 1995)

xaD − xW
LaD = = (2.5/Δ)3/2 (Fo /F1/3
z )Y
2/3
− 0.7 (16.98)
bo
and the recirculation zone (Fig. 16.45) may be described as

xs − xW
Ls = = 1.35 cos δ(f − 3), (16.99)
(bh)1/2
xes − xW
Les = = 2.5Ls , (16.100)
(bh)1/2
ys
Bs = = 6.4( cos δ)1/2 F−1
z Y
1/3
− 0.05 . (16.101)

These results apply for δ > 15◦ , and wave C in particular behaves differently for
smaller junction angles, which are irrelevant in sewer applications, however. The
upper limit of application is δ < 70◦ . According to experiments, junctions with a
right-angled branch do always choke. Provided a junction should have a thoroughly
supercritical flow, the limit junction angle should be about 60◦ . Typical junction
angles are thus 30◦ < δ < 60◦ .
16.3 Supercritical Flow 459

Fig. 16.45 Recirculation zone (bottom) and wave B (top) for junction angles δ = (a) 30◦ , (b) 60◦

The transition (subscript t) between compact and wall-type waves B occurs


approximately for

Fot = 5( sin δ)1/2 (Fz − 2)1/3 Y −1/3 . (16.102)

With these indications, the main features of supercritical junction flow are given.

Example 16.7 The approach flow conditions of a channel junction are Qo =


10 m3 s – 1 , ho = 0.8 m, bo = 1.5 m and Qz = 4 m3 s – 1 , hz = 0.45 m, bz = 1 m.
Describe the flow structure for a junction angle of δ = 35◦ !
Approach branches Fo = 10/(9.81 · 1.52 0.83 )1/2 = 2.98,
Fz = 4/(9.81 · 1.02 0.453 )1/2 = 4.23, Y = 0.8/0.45 = 1.78,
b̄ = 1.225 m, h̄ = 0.60 m,
(bh)1/2 = 0.86 m, f = 2 · 2.98 · 4.23/(2.98 + 4.23) = 3.50.
Limit conditions FLo = 2 + (35/120)4.23/1.78 = 2.69 < 2.98,
FLz = 2 + (1/8)(2.98 + 5)1.78 = 3.78 < 4.23 according to
Eqs. (16.88) and (16.89), choking does not occur.
460 16 Special Manholes

Wave C tan θ = (1/2)35◦ (π/180◦ )(4.23/2.98)2/3 /1.78 = 0.22,


thus θ = 12.1◦ ,
xMC –xP = 0.60 · cos12.1◦ 4.3(3.50 – 2) = 3.78 m
hMC = [2 + 0.25(3.50 · sin35◦ )2 ] 0.60 = 1.80 m.
Wave B With Δ = 0.55 + 0.05 · 35◦ = 2.30,
xaB –xP = [(1/2.30)(2.98/4.231/3 )1.782/3 ]1.5 = 1.77 m
xMB –xP = 1.65 · 1.77 = 2.9 m
xeB –xP = 1.35cos35◦ [2.98(4.23 · 1.78)1/3 + 5]0.86 =
10.3 m, yMB = 0.3(0.1 + sin35◦ ) 2.98 · 4.231/3 0.6 = 0.59 m
hMB = [1 + 0.25 · 3.52 ]0.60 = 2.44 m
Recirculation xs –xW = 0.86 · 1.35cos35◦ (3.5–3) = 0.5 m
zone xes –xW = 2.5 · 0.5 = 1.25 m
ys = 6.4 · 1.22(cos35◦ )1/2 [4.23–1 1.781/3 –0.05] = 1.67 m.

This indicates a relatively high wave C, and a slightly larger wave B. The
width of the recirculation zone is with ys > bo physically impossible. With
Fot = 5(sin35◦ )1/2 (4.23–2)1/3 /1.781/3 = 4.08 > Fo , wave B is compact.

16.3.7 Methods of Shockwave Reduction

The preceding discussion has demonstrated that perturbations of a supercritical flow


increase significantly with the approach Froude number Fo . Because the effects
of viscosity and wall roughness on the decay of wave amplitudes are small, and
disadvantages such as flow concentrations, surface aeration and local erosion may
occur, the reduction of shockwaves is of significance. Basically, shockwaves may
be reduced with various methods, as illustrated in Fig. 16.46 (Vischer and Hager
1994):
• Transverse bottom slope Soy to compensate the centrifugal acceleration of a
particle moving on a streamline radius R, where

Vo2
Soy = . (16.103)
gR

• Multiple vanes to reduce the total channel width b in various sub-sections


of reduced maximum wave height, an approach impractical in sewers due to
clogging, however.
16.3 Supercritical Flow 461

Fig. 16.46 Possibilities of shockwave reduction, for details see main text

• Transverse step to guide a flow with a wall minimum depth resulting in a


more horizontal surface profile. This approach is unsuitable mainly because of
abrasion.
• Cover plate to restrict the height of a wall wave.
• Application of wave interference techniques based on the superposition of two
identical waves of different sign.

Elements such as guide walls, transverse steps and application of wave interfer-
ence might be justified in hydraulic structures, but they are not a design in sewer
techniques. Therefore, only two methods of shockwave reduction in sewers are
available:

• Reduction of shock number, or


• Application of cover plates.

Shockwaves increase often linearly with the governing shock number.


Accordingly, shockwaves can be reduced in the design phase by either limiting the
approach flow Froude number, or by limiting the perturbation intensity, such as
the deflection angle, the relative curvature or the junction angle. Reducing a shock
number is thus a direct design method that adds to the performance of a sewer.
Because the Froude number of a sewer flow is F = Q/(gDh4 )1/2 and the dis-
charge is a design variable, F reduces only by increasing either the diameter D or,
more significantly, the flow depth h. If considerations refer to a long sewer reach
where uniform (subscript N) flow has nearly established, one has from Eq. (5.15) for
yN < 0.8

1/2
yN = 1.15qN (1 + qN )1/2 . (16.104)

Eliminating yN , FN can be expressed with the relative discharge qN = nQ/(S1/2 8/3


o D )
as

Q/(gD5 )1/2 0.76χ


FN = = . (16.105)
1.32qN (1 + qN ) 1 + qN
462 16 Special Manholes

Here, χ = S1/2 1/6 1/2


o D /(ng ) is the roughness characteristics according to Sect. 5.6.
If an average design value yN = 0.50 is considered, one has qN = 0.15, and
FN = (2/3)χ from Eq. (16.104). The Froude number for uniform flow is thus small
for small bottom slope, because the diameter effect is insignificant and roughness
cannot really be influenced for commercial pipe material.
The perturbation intensity of a flow includes:

• Deflection angle θ for contractions,


• Expansion ratio β = bu /bo for expansions,
• Relative centerline curvature b/R for curves, and
• Junction angle δ for junction structures.

The shockwave height depends on the product of approach flow Froude number
and the perturbation intensity. From constraints in space and cost, the shock num-
ber can often not be limited to a value of, say, below 1. Then, the only method to
reduce shockwaves is wave treatment by flow forcing. Based on a fundamental anal-
ysis of the problem in channel junctions, Schwalt (1993) considered the following
geometrical arrangements for active shock control:

• Bottom vane to separate the branch flows, resulting in a modification of the


approach flow direction to wave B,
• Bottom step to reduce the transverse surface slope, and
• Cover plate to limit the wall wave height.

Both, the bottom vane and the bottom step (Fig. 16.46) have a passive effect on
the shockwave but these elements are not flexible enough for variable approach
flow conditions. They are designed for a specific condition but either are ineffec-
tive or even perform poorer than the basic design. In addition, the elements are
prone to sediment deposition or abrasion, such that they cannot be recommended for
sewers.
The only reliable, efficient and simple method to reduce shockwaves in sewer
manholes is the cover plate (German: Deckplatte; French: Plaque de couverture).
It is an element that inhibits the rising of wall waves actively and it is mounted on
the benches of a manhole. Its prime action is to reduce the risk of choking flow in
sewers and thus the breakdown of supercritical pipe flow. The cover plate can be
inserted in junction and bend manholes, as described below. Disadvantages of cover
plates include corrosion and a certain clogging potential. An alternative is presented
in Sect. 16.5.
Figure 16.47a shows typical supercritical flow in a bend manhole, with a sub-
merged downstream sewer due to shockwave choking. Because larger shocks are
wall waves (Fig. 16.42b), they can often be reduced with a suitable cover plate.
According to Fig. 16.47b the plate can be made tip up to easen access and main-
tenance. The element can also be added to locations where shockwaves have
16.3 Supercritical Flow 463

Fig. 16.47 Supercritical approach flow to bend manhole (a) untreated and (b) treated flow

Fig. 16.48 Waves B and C at


junction manholes extended
with cover plate. (a) section,
(b) plan

caused damages. Contrary to most other elements, the plate is subjected with an
overpressure and vibrations and cavitation damage are no concern.
According to Sect. 16.3.6 wave B is always higher than wave C. Figure 16.48
relates to the cover plate of length Ld , width bd and height hd . To exclude overflow
of wave C, the height hd should always be larger than the height hMC of wave C.
The position of the cover plate must be such that there is no overflow from the
through branch, and a sufficient length guarantees a horizontal free surface towards
the downstream branch, because a wall wave is only partially reduced otherwise
(Fig. 16.49).

Fig. 16.49 Adaptation of cover plate in a junction manhole. (a) Poor and (b) appropriate design
464 16 Special Manholes

Schwalt (1993) conducted experiments in a rectangular channel and found:

• Cover plates are suitable for junction angles up to about 45◦ . For larger angles,
waves C and B have nearly the same height according to Eqs. (16.92) and (16.97),
and the element becomes inefficient.
• For larger junction angles, a combination of a bottom drop and a cover plate may
be suitable. The drop height has the order of the lateral flow depth. This design
needs further tests.

The following results refer to junction angles δ between 15◦ and 45◦ . The major
design elements of a cover plate (subscript p) are the position of beginning (sub-
script a), the end position (subscript e), the width and the height above the bottom.
According to Schwalt (1993) the beginning of the cover plate xap is at (Figs. 16.43
and 16.48)

xap − xP 0.27 3/4


= Y Fo Fz−1/3 − 1.5 (16.106)
bo sin δ
and its end xep should satisfy the condition

xep − xP 1.15 1/2


= Y Fo Fz−1/3 − 2.3. (16.107)
(bh)1/2 sin δ

Both relations are similar and contain the determining Froude number Fo /Fz1/3 . The
width of the cover plate is bp = ho + hz and its height hp above the invert should be

hp
= 2 + 0.67 sin δFz (Fo Y)1/3 . (16.108)
hz

The Froude number Fz of the lateral branch has a significant effect on hp .


Extreme wall waves occur for 2 < (Fo /Fz1/3 )Y1/3 < 7, and a cover plate yields
reductions up to 40%. Equation (16.108) also demonstrates the practical limits of
the recommended element. For large lateral flow depths, the position of the cover
plate is always large and the freeboard required is significant. Figure 16.50 refers
to identical flows without, and with a cover plate and illustrates a considerable
improvement of flow.

Example 16.8 Design a cover plate for Example 16.7.


With Fo = 2.98, Fz = 4.23, Y = 1.78, ho = 0.80 m, hz = 0.45 m,
(bh)1/2 = 0.86 m, Eqs. (16.106) and (16.107) give for start and end of the cover
plate xap –xP = [(0.27/sin35◦ )1.780.75 2.98 · 4.23–1/3 –1.5]1.5 = – 0.24 m and
xep –xP = [(1.15/sin35◦ )1.781/2 2.98 · 4.23–1/3 –2.3]0.86 = 2.26 m. Further, bp
= 0.8 + 0.45 = 1.25 m, hp = [2 + 0.67sin35◦ 4.23(2.98 · 1.78)1/3 ]0.45 = 2.2 m
16.3 Supercritical Flow 465

are width and height, such that the cover plate length is Lp = xep –xap =
2.26 + 0.24 = 2.5 m, i.e. practically equal to the downstream width.
Due to the large filling ratio, the cover plate height is slightly below hMB =
2.44 m. Because Fo (Y/Fz )1/3 = 2.25, the wave height cannot be much influ-
enced. With Fot = 5(sin35◦ )0.5 (4.23 – 2)0.33 1.78–0.33 = 4.08 according to Eq.
(16.102), wave B is of transitional type.

When applying these results to junction manholes, the Froude number of the
U–shaped profile must be determined. Often, the required length of the cover plate
is larger than the manhole length, downstream from point P . The downstream sewer
is then considered as a part of the cover plate (Fig. 16.51). According to Eq. (16.108)
the required downstream diameter can be quite large, particularly for larger junc-
tion angles δ and large Froude number Fz . Then, the only design possibilities are

Fig. 16.50 Junction flow with ho = hz = 40 mm, Fo = Fz = 8 and δ = 30◦ (a) without and (b)
with cover plate in plan (top) and section (bottom)
466 16 Special Manholes

Fig. 16.51 Cover plate in junction manhole for (a) 45◦ junction and straight lateral channel walls,
(b) 90◦ junction with curved lateral inlet (not recommended)

either a reduction of the shock number, or an increase of the downstream diame-


ter. Until today, no experiences on cover plates are available. If supercritical flow
has to be maintained across a manhole, bottom drops as mentioned in Sect. 16.2.4
should not be applied. Fig. 16.51a refers to a 45◦ junction manhole with a direct
transition to the downstream sewer, whereas Fig. 16.51b presents an unrealistic
design with a 90◦ branch channel, for which the junction flow almost certainly
will choke. A serious hydraulic design of junction manholes is essential to inhibit
submergence and pulsations, which at the limit may generate manhole overflow
(Chaps. 5 and 7).

16.4 Bend Manhole

16.4.1 Introduction
Each change of sewer direction both in plan and in the section requires a manhole. In
the following, the bend manhole (German: Krümmerschacht; French: Puits à coude)
is described for plan direction changes. Changes of bottom slope are often so small
that no additional losses have to be considered for subcritical flow. Transitional flows
in these structures are described in Chaps. 6 and 7.
The following discussion relates to sub- and supercritical flows separately, as
for junction manholes. In Sect. 16.4.4 the cover plate as recommended for junction
manholes is introduced. For subcritical flow, the results relating to pipe flow apply,
in analogy to other structures. For supercritical flow in bend manholes, a systematic
experimental program was conducted to investigate the shockwave effects, and their
consequences on air entrainment and choking of the downstream sewer.

16.4.2 Subcritical Flow


As for junction manholes, only some information on bend manholes is currently
available. For subcritical flow, the knowledge on the head loss coefficient in terms
of manhole geometry is essential to determine the backwater effect in the upstream
sewer.
16.4 Bend Manhole 467

Table 16.2 Head loss coefficients ξ k in bend manhole for geometries of Fig. 16.52 (Dick and
Marsalek 1985)

Manhole Type (a) (b) (c) (d)

Angle of deviation 22.5◦ 45◦ 90◦ 30◦ 60◦ 90◦ 90◦ 90◦

ξk 0.3 0.6 1.85 0.45 0.9 1.6 1.1 0.55

Dick and Marsalek (1985) investigated the manhole configuration with 100%
benches (Chap. 14). In the average, this design has a loss coefficient ξ k =
0.05–0.10, whereas bend manholes with 50% benches have a ξ k value of about
0.10–0.20. The poor configuration without benches at all involves ξ k values of about
0.20–0.30.
The bend (subscript k) manhole has four major designs, namely with 0, 50%
and 100% benches, and a configuration with a locally expanding U-shaped profile.
Table 16.2 summarizes some experimental observations relative to the head loss
coefficient ξ k = H/[Vo2 /2g] for pressurized manhole flow. The effect of manhole
filling was not as significant as for the simple through-flow manhole (Chap. 14). One
may note a considerable increase of ξ k with the deviation angle δ k , and a reduction
of ξ k as the bench height increases. Compared to the standard design case (c) of
Fig. 16.52, case (a) has almost the double loss, and case (b) a loss of 150%, whereas
case (d) involves a loss reduction to some 50%. The latter design is costly and not
recommended, however.
Marsalek and Greck (1988) considered bend manholes of quadratic plan with
identical diameters of the up- and downstream sewers. For larger relative submer-
gence Ss = ss /Do > 1.8, no submergence effect was noted such that the head loss
varies only with the manhole geometry. For the configuration corresponding to
Fig. 16.52c, the head loss coefficient is also ξ k = 1.1.
Contrary to the previous observations, Johnston and Volker (1990) related their
results to the upstream pipe Froude number Fo = Vo /(gDo )1/2 . In the discussion of
the results, their data were represented for δ = 90◦ and Ds /Do = 4 as

Fig. 16.52 90◦ bend manholes, geometry in transverse section (top) and plan (bottom)
468 16 Special Manholes

ξk = Ck F−1
o . (16.109)

The coefficient of proportionality Ck varies mainly with the submergence ratio


Ss and the manhole geometry. For 100% benches, the results are Ck = 0.3 for
Ss = ss /Do ∼
= 1.4, and Ck = 0.22 for Ss ∼
= 5. Therefore, ξ k decreases with increasing
submergence, in analogy to the standard manhole (Chap. 14).

Example 16.9 Given a bend manhole of deflection angle δ k = 45◦ . What is


the approach flow depth for a downstream sewer with Du = 0.70 m, hu =
0.52 m and Q = 0.3 m3 s–1 for 100% benches?
With a downstream filling ratio yu = 0.52/0.70 = 0.74, the cross-sectional
area is Fu /Du2 = yu1.4 = 0.66 from Eq. (5.16)2 , thus Fu = 0.66 · 0.72 =
0.32 m2 and Vu = 0.3/0.32 = 0.93 ms–1 . According to Table 16.2, case b)
has nearly a loss coefficient of ξ k = 0.7, and for case c) one may admit ξ k =
0.7(1.1/1.6) = 0.48 for δ k = 90◦ , i.e. Hk = 0.48 · 0.932 /19.62 = 0.021 m.
Because Fu = Q/(gDu hu4 )1/2 = 0.3/(9.81 · 0.70 · 0.524 )1/2 = 0.42 < 0.5, one
may set ho = hu + Hk = 0.54 m, because the bottom slope compensates
approximately the friction slope.

Partially-filled flow in bend manholes with δ k = 45◦ and 90◦ was investigated in
terms of head losses and transverse free surface slope. Because both the upstream
and downstream sewers were circular pipes with an U-shaped manhole of 100%
benches, the head loss Hb refers only to the manhole length, excluding the head-
loss in the downstream
  sewer due to flow realignement. The headloss coefficient
ξk = Hb / Vd2 /(2 g) related to downstream (subscript d) velocity is

ξk = Ck y−1
d , yd < 0.90 (16.110)

with yd = hd /D as the downstream sewer filling, Ck = 0.07 for δ k = 45◦ , and Ck =


0.085 for δ k = 90◦ . Accordingly, the headloss coefficient decreases as the filling
ratio increases. This effect may be explained with the deviation of the circular cross-
sectional shape as yd decreases. The range of Froude numbers tested was 0.30 <
Fo < 0.70.
The maximum superelevation hM between the outer and inner walls of a bend
manhole depends on the ratio of centrifugal to gravitational accelerations. With Ra
as the axial radius of curvature, D the width of the U-shaped profile and Vo as the
approach flow velocity, the potential vortex flow theory gives hM = DVo2 /(gRa ).
By plotting Φ = (hM /D)[gRa /Vo2 ] as a bend superelevation coefficient against the
16.4 Bend Manhole 469

approach flow filling ratio yo yields Φ = 0.80 for both δ k = 45◦ and 90◦ up to yo =
0.60, and then an increase of Φ to 1.20 for 45◦ , and a decrease of Φ = to 0.40 for
90◦ , for yo = 1.
Velocity profiles for two typical flows in 90◦ -manhole bends are shown in
Fig. 16.53. One notes an essentially uniform velocity distribution both in the stream-
wise and the radial directions, except for the boundary layers. A detailed data
analysis indicates no real velocity increase towards the curvature center, as would
follow from the potential vortex theory. One may also note from the free surface
plot that the maximum flow depth is located at ∼ = 60◦ , whereas the minimum is at

Fig. 16.53 Velocity field for various layers Z = z/D in bend manhole for subcritical flow (a) yo =
0.45, Fo = 0.41 and (b) yo = 0.86, Fo = 0.56. Flow surface at bottom
470 16 Special Manholes

75◦ downstream from the approach flow section. The two flows behave essentially
similar, yet with a more complex free surface topography due to impact on the
downstream manhole wall for yo = 0.86.
Figure 16.54 shows various typical bend flows as occur in sewage treatment
stations. In Fig. 16.54a the Froude number is estimated to 0.40 with a typical
standing surface wave pattern. The flow in a 90◦ bend may dissipate a consid-
erable amount of energy, with a transition from supercritical to subcritical flow
(Fig. 16.54b). Pump cost can thus be notable. Such flows have not yet been consid-
ered with laboratory modelling. The flow in Fig. 16.54c has an even larger approach
flow Froude number due to a small drop structure, resulting in an oblique direct
hydraulic jump.
Bend manholes are typical structures of a sewer system. Figure 16.55a refers to
a 60◦ bend manhole. Figure 16.55b shows a bend manhole during low discharge

Fig. 16.54 Bend flow in


sewage treatment stations
16.4 Bend Manhole 471

Fig. 16.55 Bend manholes


looking (a) through the
manhole shaft, (b) from
benches

condition, with a considerable sludge layer on the benches. These conditions may
produce a slippery walkway with a potential danger for accidents.

16.4.3 Supercritical Flow


Supercritical flow in bend manholes is similar to the flows discussed in Sect. 16.3.5.
The difference is mainly due to the downstream sewer limiting the flow in its vertical
extension. At the downstream manhole end, the flow may impinge onto the manhole
wall and hydraulic impact jumps may develop. In contrast, the flow in an open chan-
nel bend has no such limitations. The following is a summary of current knowledge
on bend manholes, involving also the bend cover to increase the discharge capacity
and to improve the flow pattern of the downstream sewer.
Christodoulou (1991) analyzed supercritical flow in bend manholes. Figure 16.56
shows the geometry considered, with angles δ k = 0 and 90◦ . The head loss Hs was
related to the approach flow velocity Vo . The variables are the manhole diameter
ratio δ s = Ds /D, the relative drop height Zs = zs /D, the bend deflection angle
δ k , the approach flow bottom slope So and the drop number D = Vo /(gzs )1/2 .
From observations no effect of So on the loss coefficient ξ k = Hs / [Vo2 /(2g)] was
deduced. For ξ k = 90◦ the data follow

ξk = 0.2 + 2.3D−2 . (16.111)


472 16 Special Manholes

Fig. 16.56 Supercritical flow


in bend manhole (a) plan, (b)
section

The first term 0.2 can be identified as the loss due to the manhole, and the second is
the loss due to the impact (subscript I) on the manhole, with HI = 1.15zs .
With hs as the flow depth in the manhole downstream from the drop (Fig. 16.56),
and ys = hs /zs , the data for δ k = 90◦ may be expressed as
0.3
ys = 0.85 So−1 −Som D1.5 (16.112)

with So < 10% as the bottom slope in [%] and Som = 0.15% as the minimum value.
For δ k = 0◦ , the constant 0.85 in Eq. (16.112) has to be replaced with the constant 1.
To inhibit choking of the downstream sewer, the condition hs /D < 1 must be satisfied,
and submergence into the approach flow sewer is suppressed for ys < 1.

Example 16.10 Given an approach flow sewer with So = 6%, D = 1.0 m,


1/n = 90 m1/3 s−1 and Q = 1.2 m3 s−1 discharging into a bend manhole with
δ k = 45◦ . What is the height of the drop?
For uniform approach flow the dimensionless discharge is qN =
0.011 · 1.2/0.061/2 = 0.054, thus according to Eq. (5.15) yN = 0.926[1 –
(1−3.11 · 0.054)1/2 ]1/2 = 0.276 and hN = 0.276 · 1.0 = 0.28 m. Further, with
Fo /Do2 = y1.4
N = 0.276
1.4 = 0.165 from Eq. (5.18) , the cross-sectional area is
2
Fo = 0.165 · 1 = 0.165 m2 such that Vo = Qo /Fo = 1.2/0.165 = 7.3 ms–1 .
Starting from Eq. (16.112), the resulting flow depth in the manhole is
ys = 0.9(6−1 −0.15)0.3 [Vo /(gzs )1/2 ]1.5 = 0.94/zs0.75 . If submergence has
to be inhibited, i.e. ys < 1, then zs = (0.94)−4/3 = 1.09 m, correspond-
ing to D = 7.3/(9.81 · 1.09)1/2 = 2.23. If the manhole should flow free,
then hs /D = 0.9(6−1 −0.15)0.3 [Vo /(gzs )1/2 ]1.5 (zs /D) ≤ 1, with the solution
zs = 1.071/4 = 1.02 m. The drop height should thus be about 1 m with the
manhole so long that the jet does not impact the downstream manhole wall
(Chap. 11).
16.4 Bend Manhole 473

Del Giudice et al. (2000) considered bend manholes with a supercritical approach
flow for deflection angles 45◦ and 90◦ . The effect of bottom slope can be neglected
provided So < 2%. The bend manholes tested had a relative curvature ρ a = D/Ra =
1/3, with Ra as the radius of the bend axis. The benches were 150% of the diameter
because of space limitations with the lateral manhole extension.

Untreated Manhole Flow


Figure 16.57 relates to the 90◦ bend manhole with an approach flow Froude num-
ber Fo = 2.18 and an approach flow filling yo = 0.35. A view from the approach
flow sewer shows a significant but smooth increase of the outer wall profile and a

Fig. 16.57 Photographs of supercritical flow in bend manholes, for details see text
474 16 Special Manholes

decrease of the inner wall profile (Fig. 16.57a). The complete free surface along the
manhole is shown in Fig. 16.57b, with the shockwave clearly separated and nearly
vertical along the curved channel. At the end of the manhole, the flow impinges
on the manhole endwall, with a swell of maximum height zs at the outer wall
(Fig. 16.57c). A swell portion falls back on the supercritical flow which may choke
if the recirculation zone is too long. The capacity of the manhole is thus governed
by the features of the swell. For large discharges, the recirculation can even have an
influence on the inner wall profile, as is seen in the photograph. The curved shock
front as seen towards the manhole upstream end is also shown in Fig. 16.57d. Again,
the free surface gradients are seen to be immense as compared to usual hydraulic
configurations.
Figure 16.58a and b relate to a complete overview, including the approach
pipe flow, the outer wall wave and impingement onto the manhole end wall.
Figure 16.58c shows the flow in the downstream sewer with the swell on the left
side, contraction of flow due to the pipe inlet and formation of wave 1 on the
inner side of the downstream sewer. Wave 2 on the outer side can be seen in its
beginning. In all cases tested, with always a free pipe outflow, these waves were
never a problem, and choking never occurred in the downstream pipe due to wave
formation.
The velocity field pertaining to the flows presented in Figs. 16.57 and 16.58 is
shown in Fig. 16.59 for various levels Z = z/D above the invert, with z as the

Fig. 16.58 Manhole flow for supercritical approach flow, for details see text
16.4 Bend Manhole 475

Fig. 16.59 Velocity field


pertaining to experimental
run shown in Figs. 16.57 and
16.58. Bottom: average
velocities (left) and free
surface (right)

vertical coordinate. Comments as for subcritical bend flow apply: (1) In a cer-
tain height above the invert, the variation of velocity is small, except close to
the walls and in the recirculation region; (2) The transverse velocity has a slight
tendency to increase towards the center of curvature: (3) Except for the bottom
layer, the velocity decreases with height; (4) The effect of shock wave may not
really be detected in the velocity field but it clearly follows from the free surface
plot. Accordingly, supercritical flows across bend manholes can be approximated
with an almost constant velocity equal to the average approach flow velocity,
and the particular hydraulic concern is the free surface because of extreme wave
formation.
The extreme flow depths have a maximum (subscript M) flow depth hM and loca-
tion θ M at the outer wall, and a minimum (subscript m) flow depth hm at location θ m ,
with θ as the angle measured along the manhole from the intake section (Fig. 16.60).
Based on Sect. 16.3.5, the data for the extreme wave elevations ZM = YM 1/2 − 1
476 16 Special Manholes

Fig. 16.60 Definition plot for supercritical flow across bend manhole (a) plan, (b) section

and Zm = Ym1/2 − 1 may be expressed with the approach flow bend number
Bo = ρα1/2 Fo as

ZM = Zm = 0.50B2o . (16.113)

The difference in ZM between the rectangular and the U-shaped channels is obvi-
ously a shape effect. Note that Fo = Vo /(gho )1/2 , because the upper portion of the
U-shaped channel is rectangular.
The locations of extreme wave heights vary with ρ a Fo , as for the rectangular
channel bend. The results for Bo < 1.50 are

tan θM = 2.8(ρa Fo )2 , (16.114)



tan θm = 2(ρa Fo ). (16.115)

The wave profile along the outer bend wall can also be described with Eq. (16.82).
The height of the swell Zs = zs /D is equal to

Zs = σs B2o , Bo < 1.5 (16.116)

with σ s = 0.80 for δ k = 45◦ and σ s = 0.50 for δ k = 90◦ . The swell may be higher
or smaller than the maximum wave height, depending on its location, therefore.
Waves 1 and 2 in the downstream sewer can be described by their distances d1
and d2 from the manhole end, and their respective heights h1 and h2 . The details
16.4 Bend Manhole 477

are given by Del Giudice et al. (2000). As already mentioned, these waves have
never had an effect on the flow in the bend manhole, and are not of design concern,
therefore.
The discharge capacity (subscript C) of a bend manhole is reached when Zs
has the order of 1.5. Then, the impact of the supercritical flow onto the manhole
end wall is so strong that the recirculating flow may choke, and a hydraulic jump
forms in the manhole. For δ k = 45◦ , the absolute maximum (subscript L) approach
flow filling was yoL = 0.70, whereas it amounted to only yoL = 0.55 for δ k =
90◦ . For yo > yoL , the manhole flow chokes, with a hydraulic jump moving into
the upstream sewer. Sewers with an approach flow filling ratio larger than either
55%, or 70%, respectively, choke and supercritical flow may not be maintained.
This finding is important in terms of manhole choking, breakdown of supercriti-
cal flow, upsurging of manhole water level up to blasting of manhole covers. The
conventional sewer design involving the full flow concept cannot be maintained
for supercritical flow, therefore. A novel design of such configurations is presented
below.
For 0.20 < yo < yoL the capacity Froude number FoC can be expressed as

FoC = 3 sin δk (1 − yo ) + yo . (16.117)

This transitional condition corresponds to a lower bound of the approach flow


Froude number, below which a supercritical manhole flow cannot be maintained.
The details of this analysis are also presented by Del Giudice et al. (2000), and it
can be stated here that supercritical flow breaks down for Fo < 1.50. Then, an undu-
lar hydraulic jump may establish first in the downstream pipe, and eventually move
upstream when decreasing Fo toward 1. As a general consequence, approach flows
with 0.75 < Fo < 1.50 should be inhibited because of the formation of a transitional
flow character. Further results relating to the energy loss across a bend manhole
for supercritical flow, and the shockfront development are not discussed here (Del
Giudice et al. 2000).

Example 16.11 Given a bend manhole with δ k = 90◦ and D = 0.60 m,


approach flow bottom slope So = 3% and roughness coefficient 1/n = 85
m1/3 s−1 . Determine the bend flow characteristics for a discharge of Q = 0.90
m3 s−1 .
Assuming uniform approach flow gives with Eq. (5.15) ho /D = 0.65, thus
ho = 0.65 · 0.60 = 0.39 m, Vo = 4.65 ms−1 , Fo = 4.65/(9.81 · 0.39)1/2 = 2.38
and Bo = 2.38(1/3)1/2 = 1.37. Therefore, hM /ho = (1+0.50 · 1.372 )2 = 3.77,
tan θ M = 2.8(2.38/3)2 = 1.76 and Zs = 0.5 · 1.372 = 0.94 from Eqs. (16.113),
(16.114) and (16.116). The characteristics of the bend flow are, therefore,
hM = 3.77 · 0.39 = 1.47 m, θ M = 60◦ and zs = 0.94 · 0.60 = 0.56 m.
478 16 Special Manholes

16.4.4 Shockwave Reduction


Abrupt Wall Deflection
As discussed in Sect. 16.3.7, wall waves may effectively be reduced with the cover
plate. Schwalt (1993) considered the mitre bend for deflection angles δ = 30◦ and
60◦ . Because bend manholes are short, his experimental indications are relevant.
For 20◦ < δ < 60◦ , the beginning xad of the cover plate is at location xP , i.e. at the
extension of the lateral wall to the opposite side (Fig. 16.48b). For a junction where
the through branch has no discharge, the position of beginning is xad = xP , and the
cover plate starts at the junction point P.
The end of cover plate is, independent of δ,

xed − xP
= 0.6F1/2
z . (16.118)
bz

The height hd of the cover plate varies with the deflection angle δ, the approach flow
Froude number Fz and the width ratio bz /bd as

hd /hz = 2 + 0.23 sin δFz (bz /bd )1/2 . (16.119)

The width of the cover plate should be equal to bd = 1.5hz for δ = 30◦ and bd = 2hz
for δ = 60◦ . Figure 16.61 shows the flow deflection for δ = 30◦ with and with-
out a cover plate. The effect of the plate is significant, and the downstream flow is
much improved, mainly as regards uniformity and shockwave damping. To further
improve the flow, Schwalt suggested a second cover plate on the opposite side.

Example 16.12 Given an abrupt channel deflection with δ = 30◦ , Fz = 6,


hz = 0.4 m and bz = bu = 0.9 m. What is the geometry of the cover plate
required?
With x = 0 at point P , the plate starts at xap = xP = 0 and its width
is bp = 1.5hz = 0.6 m. With Eq. (16.118) the end of the cover plate is at
xep −xP = 0.6 · 61/2 0.9 = 1.32 m. The cover plate height is from Eq. (16.119)
hp = [2 + 0.23sin30◦ 6(0.9/0.6)1/2 ]0.4 = 1.15 m. The plate dimensions are thus
Lp = 1.32 m, bp = 0.6 m, and hp = 1.15 m. The downstream channel should
be at least 1.15/0.9 = 1.25 m high, therefore. Because the wave height in the
channel without a cover plate would be hMB /hz = 1 + (sin δ/4)Fz2 = 5.5, i.e.
hMB = 5.5 · 0.4 = 2.2 m, the reduction of wave height is with more than 40%
significant.
16.4 Bend Manhole 479

Fig. 16.61 Deflection flow


(a) without and (b) with cover
plate in plan (top) and section
(bottom)

Bend Manhole
Shockwaves in a bend manhole with an U-shaped profile may get high, for a large
bend number Bo and a typical design approach flow filling of yo ∼ = 0.50. Also, the
supercritical flow in a bend manhole may break down if yo > yoL , as described pre-
viously. Out of the two methods to reduce shockwaves, i.e. (1) Reduction of bend
number, and (2) Cover plate, only the latter method can be often applied because
of limitations in the upstream sewer. Figure 16.62 shows a definition sketch of the

Fig. 16.62 Definition plot


for cover plate in bend
manhole (a) plan, (b) section
480 16 Special Manholes

bend manhole with a cover (subscript c), including the angles α c and β c where the
flow attaches the cover, the air discharge Qa and the downstream length La measured
from the manhole end where free surface flow is reestablished.
The optimum height of the cover plate above the manhole invert was found to be
0.90 D, providing both a large flow section below the cover, and a sufficiently large
space above it for aeration of the downstream sewer. The cover impinging angles
vary essentially with yo = ho /D and Fo = Vo /(gho )1/2 as (Del Giudice et al. 2000)

αc yo = 10(Fo − 1)−1/3 , (16.120)

βc yo = 8.5[1 + 2 exp ( − (Fo − 1)2 )]. (16.121)

If α c and β c are smaller than the manhole deflection angle δ k , the latter has no
effect on flow reattachment. To simplify the design, the complete manhole can
be covered. The cover plate should be secured against uplift pressure. A conser-
vative design involves the maximum pressure over the cover cross-sectional area
Ac = (δ k − α c )Ra D. Using the expression for the maximum wave height hM from
Eq. (16.113) gives Pc = ρg(hM – ho )Ac for the maximum pressure force. A cover
should be easily removable from the U-shaped profile for maintenance and sewer
inspection. The cover may not be suited for shockwave control in all cases of
practice.
More practical information has to be collected for the optimum arrangement of
the cover plate, and its potential for clogging a sewer. Figure 16.63 shows photos
from the lab model for δ k = 45◦ and yo = 0.49, Fo = 3.40. It can be seen that the
transition from free surface to covered flow occurs within a short reach, and that the
flow in the tailwater pipe may become annular.
The air supply Qa required for a flow without underpressure in the downstream
sewer was also determined. The air supply depends mainly on the approach flow
Froude number Fo , the water discharge Q and the approach flow filling ratio yo
because of the limited air cross-sectional area above the air-water mixture flow.
From experiments, the relative air discharge B = (Qa /Q)yo4/3 is close to zero for Fo <
2, increases steeply with Fo and remains almost constant at B = 0.24 for Fo > 4. The

Fig. 16.63 Flow below


manhole cover. (a) Plan view
with flow from left, (b) side
view of tailwater pipe
including in flow direction
cover (left) and annular pipe
flow (right)
16.4 Bend Manhole 481

experimental data follow the relation (Del Giudice et al. 2000)

B = 0.24[ tanh (Fo − 2)2 ]2 . (16.122)

Depending on yo and Fo , a bend cover may induce either free surface or full-
pipe air-water flow in the downstream sewer. Because of limited discharge, a fully
choked flow under the cover with a hydraulic jump in the upstream sewer was never
attained for both δ K = 45◦ and 90◦ . The maximum approach flow Froude num-
ber in the experiments conducted was FoM = 1 + 1.3yo–1 , and this value may give a
guidance of the upper limit conditions for the bend manhole with the cover plate.
For yo = 0.50, say, the maximum approach flow Froude number would thus be
FoM = 3.60.
The increase of discharge capacity due to the presence of the bend cover is typ-
ically 60% for yo > 0.30 and slightly less for smaller values of yo . Accordingly, the
bend cover can be regarded as an effective device to increase discharge, to improve
the flow in the downstream sewer, to inhibit flow choking and the transition to pres-
surized sewer flow. Because pressure on the cover is always positive, cavitation is
of no concern.
For both large values of yo and Fo , the downstream sewer may run full with
an air-water mixture flow. The length La of the mixture flow measured from the
manhole end is (Del Giudice et al. 2000)

La /D = 2yo (Fo − 1)2 . (16.123)

For all flows tested up to Fo = 5 such a transition to full-pipe mixture flow exhib-
ited no problems, provided that sufficient air supply discharge Qa is available.
Accordingly, the flow downstream of the bend manhole was free of vibrations,
pulsating mixture flow, and large air pockets.
Clearly, the conditions tested in the experimental setup involved exclusively free
downstream conditions, i.e. the downstream sewer should never be submerged. A
submerged bend manhole could perform adversely because of the breakdown of the
air transport into the downstream sewer. Conditions for downstream submergence
have not yet been investigated.
The cover plate was suggested for practical application by ATV (1996).

Example 16.13 What are the characteristics of a covered bend manhole


assuming the conditions of Example 16.11?
The cover angles are from Eq. (16.120) α c = 10/[0.65(2.38 – 1)1/3 ] = 14◦
and from Eq. (16.121) β c = (8.5/0.65)[1 + 2exp(−(2.38 − 1)2 )] = 17◦ . Using
hM = 1.37 m as the maximum pressure head, the pressure force on the cover
is Pc = ρg(hM − ho )(δ k − α c )(π/180◦ )Ra D = 1.40 t. The air discharge is
482 16 Special Manholes

Qa = 0.24[tanh(Fo − 2)2 ]2 Q/yo4.3 = 0.24[tanh0.342 ]2 0.90/0.654/3 = 0.008


m3 s–1 from Eq. (16.122), i.e. no additional air is required. The length of
mixture flow is La = 2ho (Fo −1)2 = 2 · 0.39(2.38 − 1)2 = 1.49 m from
Eq. (16.123). The limit Froude number for the bend manhole is FoM = 1 + 1.3/
0.65 = 3 > Fo = 2.38 and the design is not prone to breakdown of flow.

16.5 Definite Manhole Design

16.5.1 Introduction
Supercritical manhole flow is governed by either shockwaves generated at each flow
discontinuity, or hydraulic jumps, if the discharge capacity is too small to convey a
fully supercritical flow. Whereas shockwaves involve mainly a medium increase of
the flow depth beyond a shock front, a hydraulic jump may result in the collapse of
the supercritical flow regime and a backwater effect. The latter may be considered
a serious problem for a sewer because of an abrupt change from free surface to
pressurized two-phase flow. This choking phenomenon is accompanied further with
water hammer, a decrease of the discharge capacity finally resulting in so called
geysering of wastewater out from the manhole onto public space (Fig. 16.64). Sewer
breakdown must be avoided in any case (ATV 1996, 2000).
The following intends to present the definite recommendation for the through-
flow, the bend and the junction manholes, based on extensive hydraulic modeling
at VAW, ETH Zurich. This research was conducted after it had been observed that
the cover plate does not satisfy all requirements, mainly relating to sewer access
and the potential of clogging, as already mentioned in Sect. 16.4. The purpose of
this sub-chapter is to present a method that is simple in design and allows for a
straightforward determination of all pertinent parameters.

Fig. 16.64 Geysering of


manhole in a combined sewer
(Hager and Gisonni 2005)
16.5 Definite Manhole Design 483

16.5.2 Through-Flow Manhole


A through-flow manhole is the simplest sewer manhole arrangement for control and
maintenance purposes (Chap. 14). The manhole of U-shaped flow profile and length
L is connected to equal up- and downstream sewers of diameter D. Figure 16.65
shows a definition sketch involving the approach flow depth ho and velocity Vo . For
yo = ho /D ≤ 0.50, the flow remains entirely in a circular-shaped pipe, whereas the
flow abruptly expands at the manhole inlet for yo > 0.50, forming a side depression
(Sect. 16.3.4) which is followed by a shockwave of height hi shortly downstream
because of flow impact onto the side walls. Whereas this phenomenon is relatively
small, a more dramatic change occurs at the manhole outlet because of flow impact
onto the upper portion of the circular profile, resulting in a shaft (subscript s) flow
depth hs . Depending on its height relative to D, the flow may either continue as a
supercritical flow, or it breaks down due to the formation of an impact hydraulic
jump. Choking then results at the manhole outlet because the jump formation and
the breakdown of the air transport from the up- to the downstream sewer reaches
(Fig. 16.66). If the discharge increases fast, the choking phenomenon may initiate
even geysering, as previously described.
Given that the U-shaped profile corresponds essentially to a rectangular channel,
the determining Froude number is FU = Q/(gD2 ho3 )1/2 . The relative shaft outflow
depth was experimentally found to (Gargano and Hager 2002)

hs /ho = 1 + (1/3)(FU yo )2 . (16.124)

Therefore, the relative wave amplitude [(hs – ho )/ho ] increases quadratically with
FU yo , or the ratio [(hs – ho )/D] depends exclusively on FU .
The discharge capacity (subscript C) QC of this manhole is of design interest.
According to Eq. (16.124) the approach flow filling yo is relevant. The transition
from free surface to pressurized manhole flow may be accounted for by the capacity
Froude number FC = QC /(gD5 )1/2 . Gargano and Hager (2002) proposed for 0.70 <
yo < 0.75

FC = 14.6 − 17.3yo . (16.125)

Fig. 16.65 Hydraulics and


design of through-flow
manhole (a) section, (b) plan
(Hager and Gisonni 2005)
484 16 Special Manholes

During all tests, no free surface flow resulted if yo > 0.75, but choking never occurred
for yo < 0.70. In the average, the choking Froude number amounted to FC = 2.
The current sewer design practice accounts for the so-called full-flow approach
(Chap. 5), involving a relative sewer filling of some 85%, independent of the
flow conditions. This condition was originally introduced for nearly uniform flows,
which differ significantly from the supercritical flows previously described. The
observations presented clearly indicate that the standard design procedure results
always in a breakdown of the manhole flow. Accordingly, supercritical flows in
through-flow manholes must be limited both in the filling ratio and the discharge
capacity, to ensure no change of the flow regime.

16.5.3 Bend Manhole


The bend manhole may be often found in the urban infrastructure, given that roads
are normally arranged in a rectangular grid. Of particular interest is the 90◦ bend
manhole, but also the 45◦ deflection angles may be relevant. The average bend radius
is usually Ra = 3D, with D as the sewer diameter. One might think that the 90◦
bend manhole is more critical in terms of discharge capacity than the 45◦ manhole.
Figure 16.67 shows a definition sketch involving the approach flow depth ho and
velocity Vo for a deflection angle of δ = 45◦ . As discussed in Sects. 16.2 and 16.4.3,
two shockwaves form along the inner and the outer walls. In the following, only the
wave along the outer wall of maximum height hM is considered. Del Giudice et al.
(2000) found with FU = Q/(gD2 ho3 )1/2 that

hM /ho = [1 + 0.50(D/Ra )FU 2 ]2 . (16.126)

Fig. 16.66 Choking flow at


through-flow manhole outlet
for yo = 0.75 and Fo = 1.30,
(a) section, (b) view from
upstream, (c) impact flow
(Hager and Gisonni 2005)
16.5 Definite Manhole Design 485

Fig. 16.67 Bend manhole


with manhole extension (a)
plan where S is swell, (b)
section (Hager and Gisonni
2005)

The angle θM of maximum wave location is located between 35◦ and 55◦ measured
from the manhole inlet (Gisonni and Hager 2002a). The discharge capacity of the
45◦ bend manhole as compared to a 90◦ deflection is therefore dramatically reduced
due to the presence of the maximum wave at the manhole outlet.
To improve the capacity of this manhole, a straight tailwater manhole extension
of length 2D was added to the structure, as shown in Fig. 16.67. The extension length
was fround from detailed hydraulic tests, resulting in a wave maximum upstream of
the manhole outlet and a second wave maximum within the tailwater sewer which
does not result in flow choking. The manhole extension increases significantly the
discharge capacity, which was determined for yo < 2/3 from model tests to (Gisonni
and Hager 2002a)

FC = (3 − 2yo )y3/2
o . (16.127)

The discharge capacity of bend manholes is thus significantly smaller than of the
corresponding through-flow manhole, with a maximum of FCM = 0.90 for yo =
0.67, and only FC = 0.80 for a typical sewer filling of yo = 0.60. Note that in all
tests, the flow across the bend manhole choked if the approach flow filling was in
excess of 65%, as compared to 75% for the through-flow manhole.
Figure 16.68 shows typical flow features in a bend manhole prior to flow chok-
ing. The discharge capacity may be increased if the tailwater sewer diameter Dd is
increased thereby using a manhole extension length of 2Dd instead of 2D. No tests
were so far conducted to analyze the effect of an increased tailwater diameter on
both the approach flow filling and the discharge capacity. Note also that the choking
was always related to so-called gate flow type, as described by Hager (1994).

You might also like