You are on page 1of 11

Urban Water Journal

ISSN: 1573-062X (Print) 1744-9006 (Online) Journal homepage: www.tandfonline.com/journals/nurw20

Hydraulic design aspects for supercritical flow in


vortex drop shafts

G. Crispino, M. Pfister & C. Gisonni

To cite this article: G. Crispino, M. Pfister & C. Gisonni (2019) Hydraulic design aspects
for supercritical flow in vortex drop shafts, Urban Water Journal, 16:3, 225-234, DOI:
10.1080/1573062X.2019.1648531

To link to this article: https://doi.org/10.1080/1573062X.2019.1648531

Published online: 30 Jul 2019.

Submit your article to this journal

Article views: 253

View related articles

View Crossmark data

Citing articles: 9 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=nurw20
URBAN WATER JOURNAL
2019, VOL. 16, NO. 3, 225–234
https://doi.org/10.1080/1573062X.2019.1648531

RESEARCH ARTICLE

Hydraulic design aspects for supercritical flow in vortex drop shafts


a
G. Crispino , M. Pfisterb and C. Gisonni a

a
Department of Engineering, Università della Campania Luigi Vanvitelli, Aversa, Italy; bDepartment of Civil Engineering, Haute Ecole d’Ingénierie et
d’Architecture (Haute école spécialisée de Suisse occidentale, HES-SO), Fribourg, Switzerland

ABSTRACT ARTICLE HISTORY


Vortex drop shafts, as special sewer manholes, operate optimally if an adequate energy dissipation is Received 20 February 2019
guaranteed and the integrity of the structural components is safeguarded. The results of an experi- Accepted 18 July 2019
mental study on a vortex drop shaft with supercritical inflow are discussed herein. The hydraulic KEYWORDS
behaviour of the spiral inlet, the vertical shaft and the dissipation chamber is described. Based on Dissipation chamber; energy
detailed flow observations, useful recommendations for designing these structures are provided. It is dissipation; physical model;
demonstrated that a relation adopted for supercritical bend flows provides a reliable estimation of the sewer manhole; supercritical
maximum wave height along the inlet. A procedure for predicting the rotational flow angles and the flow; vortex drop shaft
velocity distribution along vertical shafts with swirling flows is developed. Water levels and pressure
measurements in the dissipation chamber are further analysed to identify maximum forces acting on
the chamber invert and to derive preliminary design equations.

Introduction predict shock wave heights occurring along the inlet struc-
ture were provided by Hager (1990). Del Giudice, Gisonni,
Drop manholes are common elements in urban drainage
and Rasulo (2010) coupled a conventional spiral inlet for
systems. The magnitude of the drop height strictly relates
subcritical approach flow with an upstream negative step to
to the specific purpose for which a drop manhole is needed.
enhance the structural performance. The hydraulics of the
Often, several sewer collectors arrive at different elevations
vertical shaft, along which an annular flow (Jain 1987) falls
at a node of the drainage system, requesting a simple drop
down to the outlet structure, was documented by
structure. Other ordinary drop manholes are adopted to
Kellenberger (1988). Camino, Zhu, and Rajaratnam (2014)
limit sewer collector slopes and, consequently, flow veloci-
studied flow regimes and velocities and air distributions at
ties and flow aeration. Under these circumstances, the mis-
different fall heights along the vertical shaft. Measurements
alignment of collector bottoms gives small drops, which are
of wall pressures and flow thickness in the shaft were
typically inferior to the characteristic dimension (the dia-
reported by Zhao et al. (2006). A dissipation chamber at
meter, for sewer pipes) of the involved collectors (De
the toe of the vertical shaft is foreseen to reconvert vertical
Martino, Gisonni, and Hager 2002). In other cases, the
into horizontal flow and to de-aerate the flow. Dimensions
downstream collectors, in which upstream sewers convey
for this structure were recommended by Hager and
respective discharges, are located at a significantly lower
Kellenberger (1987). Del Giudice, Gisonni, and Rasulo
elevation for sanitary and technical requirements. This is
(2008) proposed non-standard dissipation chamber config-
an everyday problem for hydraulic engineers, who have to
urations to optimize the hydraulic behaviour of the outflow,
face the difficulty to consider manholes with drop heights
entering the downstream collector.
up to tens of meters. Then, Vortex Drop Shafts (VDSs) can
An analysis of the reported investigations directs attention
be applied to connect up- and downstream sewer collectors
to some concerns on the hydraulic behaviour of supercritical
(Jain and Kennedy 1983; Vischer and Hager 1995).
VDSs components, to be further developed:
Usually, VDSs are designed following guidelines and
empirical recommendations from the literature. Early com-
(1) The inlet structure dimensions at the top of the
prehensive studies on VDS hydraulics were conducted by
vertical shaft are often quantified following the
Drioli (1969), Knauss (1987), Kellenberger (1988), Quick
empirical results of Kellenberger (1988) and Hager
(1990) and Motzet and Valentin (2002) for sub- and super-
(1990) for supercritical approach flows. The freeboard
critical approach flows. Starting from these benchmarks,
is related to the expected maximum shock wave
different experimental investigations succeeded. Jain and
height. The latter may be estimated applying the
Kennedy (1983), Jain (1984), Hager (1990) and Yu and Lee
design relation provided by Hager (1990). New
(2009) focused on the inlet structure guiding the approach
experimental data would underline its relevance
flow into the shaft. Both the standard geometry and the
and, eventually, validate its accuracy for different
hydraulic performance were detailed, depending on the
inlet structures.
approach flow energy content. Empirical equations to

CONTACT G. Crispino gaetano.crispino@unicampania.it


© 2019 Informa UK Limited, trading as Taylor & Francis Group
226 G. CRISPINO ET AL.

(2) Open questions relating to the vertical shaft perfor-


mance persist. For instance, annular flow rotational
angles and flow velocities along the vertical shaft can
be specified. An analytical calculation of the flow thick-
ness, angles and velocities was proposed by Liu et al.
(2018). However, this procedure was applied for a new
type of VDS with a special inlet structure, so that
a verification of the Kellenberger (1988) geometry
remains open.
(3) The operation of the dissipation chamber represents
another crucial issue. Energy dissipation, flow de-
aeration and free-surface oscillations of the flow issued
in the downstream collector are hydraulic features to
be specifically considered in the chamber design, espe-
cially under overload scenarios.

The present research focuses on the above-mentioned con-


cerns, working with a physical large-scaled VDS model. The
model represents a ‘traditional’ VDS, as described by Hager
(2010) and Gisonni and Hager (2012), with the unique specifi-
city to present a junction chamber collecting flows from four
Figure 1. (a) Definition sketch (■ for pressure measurement devices, → for the
sewer collectors upstream of the inlet structure. Pfister et al. flow direction, – - – for free-surface) and (b) photographs of the model.
(2018) demonstrated that this peculiarity does not affect the
overall performance of the VDS. The physical model data were
collected by Crispino et al. (2016) and Pfister et al. (2018). They Table 1. Test program (prototype scale).
are analysed in the following to explore the hydraulic beha- Test Q (m3/s) ho (m) Fo (-) FC (-)
viour of the VDS model, from the inlet structure to the dis- A 1.17 0.07 9.23 0.04
sipation chamber. The outcomes of the present research can B 2.53 0.16 6.29 0.09
be usefully extended to any VDS layout with an inlet structure, C 6.43 0.23 4.61 0.23
D 9.54 0.59 3.36 0.34
conveying an annular swirling flow in the vertical shaft, and E (design discharge) 12.74 0.87 2.50 0.46
a dissipation chamber. F (overload) 14.54 0.99 2.37 0.52
G (overload) 15.77 1.04 2.36 0.56
H (overload) 17.39 1.11 2.30 0.62
Physical model
The physical VDS model was built at the Laboratory of of ±1%. The reduced dimensions of the propeller head sen-
Hydraulic Structures (LCH) of École Polytechnique Fédérale sor (ø22∙28 mm) led to disturb the flow marginally. Flow
de Lausanne (EPFL). The model replicated a future prototype depths in the chamber were visualized and measured by
serving the city of Cossonay, Switzerland, with a geometrical measuring strips on the transparent walls. Piezometric and
scale factor of 1:7.82. As scaled, the model was sufficiently dynamic pressures at the chamber invert were collected by
large to limit scale effects, especially with reference to the air- using six piezometers and dynamic pressure transmitters,
water mixture flows (Pfister and Chanson 2014). Figure 1 gives respectively, installed at x/Lch = 0.01, 0.12, 0.24, 0.48, 0.72
an overview of the model. In the following, the VDS is and 0.96 (Figure 1(a)). The dynamic pressure transmitters
described with prototype dimensions. (Keller Series 25, Keller AG für Druckmesstechnik,
The inlet structure was made of PVC. It was shaped as Winterthur, Switzerland) had a full-scale accuracy of ±0.5%.
a spiral according to Kellenberger (1988), with eccentricities They were connected to a 16-bit data acquisition board. The
e1 = 0.20 m and e2 = 0.80 m relative to the shaft axis. The first sampling frequency was 1.0 kHz and 60,000 samples were
spiral radius was R1 = 3.10 m. The vertical shaft and the taken per each test-run. Downstream of the chamber,
dissipation chamber were made of transparent PVC. The a tailwater tunnel conveyed the flow into the laboratory
shaft had a diameter of Ds = 2.40 m and a vertical height of recirculation system. Discharges were monitored by inductive
Ls = 59.30 m. The dissipation chamber was designed according flowmeters with a full-scale accuracy of ±0.5%.
to Hager and Kellenberger (1987). It was a parallelepiped, with In total, 8 test-runs were performed, all of which under super-
a length of Lch = 10.00 m, a width of Bch = 3.00 m and a height critical approach flow with Froude numbers Fo = Q/(g∙b2ho3)0.5
of Hch = 5.00 m. The chamber was equipped with a vertical de- ranging between 2.30 and 9.23. Here, Q is the discharge, b is the
aeration pipe of 1.40 m diameter. width of the rectangular inlet channel upstream of the spiral inlet
Flow depths along the spiral inlet were measured with and ho is the flow depth at the spiral entrance. The capacity
a point gauge connected to an electronic contact-time indi- Froude number relative to the shaft was calculated as FC
cator, with a reading accuracy of 0.5 mm. Flow velocities = Q/(g∙Ds5)0.5 and it ranged between 0.04 and 0.62. The test
were detected by a propeller type Schiltknecht MiniAir® 20, program, also indicating design and overload discharges, is
with a measuring range from 0.7 to 20.0 m/s and an accuracy given in Table 1.
URBAN WATER JOURNAL 227

Spiral inlet
The spiral inlet forces the approach flow to rotate, to trans-
form from rectilinear to annular flow and to orient from hor-
izontal to vertical. For supercritical approach flows, shock
waves are generated due to the flow deflection. As observed
during the tests, the free-surface flow along the spiral inlet
occupied partially the transverse cross-section, with one or
two local standing shock waves along the outer spiral wall.
In the test represented in Figure 2 (test G), for instance, only
one shock wave occurred.
Hager (1990) observed one standing shock wave only for
highly supercritical approach flow (4.50 < Fo < 10.00) whereas
a second wave was observed for small inlet channel bottom slopes
and reduced Fo. As observed in the present investigation, the free-
surface profile along the spiral inlet was characterized by one
wave maximum, whereas a second local shock wave was present
in the spiral for large Froude numbers. The observed non-
dimensional profiles y(X) along the spiral inlet are shown in
Figure 3, where y = (h − ho)/(hM − ho), with hM as the maximum
wave height, and X = x/R1. The maximum wave amplitude y = 1.0
(h = hM) is located at X ≈ 0.3 corresponding to a spiral angle of Figure 3. Non-dimensional profile y(X) along the spiral inlet (the break line – - –
at y = 0.0 scales positive and negative semi-axes, dotted rectangles indicate
θ = 135° (Figure 1). Downstream of X = 0.3, the free-surface profile datapoints corresponding to minimum hm and maximum hM wave heights).
is decreasing with undulations for 2.30 ≤ Fo < 4.61 (tests from C to
H as shown in Table 1). A minimum was achieved for Fo = 9.23 (test
A) at X ≈ 0.4 and for Fo = 6.29 (test B) at X ≈ 0.5, instead. Downstream of this local minimum, a second wave amplitude
with −1.00 < y < −2.00 was observed for Fo = 6.29 and Fo = 9.23.
The y-value corresponding to the first shock wave is of
design interest because it dictates the height of the spiral
inlet chamber (plus freeboard). Hager (1990) estimated the
relative maximum wave height hM/R1 as a function of the
hydraulic features of the supercritical approach flow in terms
of ho, Fo, the bottom slope So, b and R1, as follows
hM h 0:5 i
¼ ð1:1 þ 0:15Fo Þ 2bho 2 R1 3 Fo  0:5So (1)
R1
Alternatively, Equation (2) provides an approximation of hM with-
out knowing the detailed inlet channel geometry. Equation (2) is
based on the present experimental data and on those of
Kellenberger (1988). It was inspired by the recommendations of
Reinauer and Hager (1997) for supercritical bend flows, stating
that the wave maximum depends on the bend number B = Fo
(b/R1)0.5. Figure 4(a) compares the normalized experimental wave
height maxima hM/ho to the bend number as
hM
¼ 2:3  B (2)
ho
As indicated in Figure 4(a), the agreement is satisfactory. Shock
wave maxima, as observed in the present VDS model, are slightly
larger than those of Kellenberger (1988). Equation (2) provides an
estimation of the relative maximum wave height hM/R1 with
R2 = 0.85, being larger than the determination of Equation (1)
(R2 = 0.80) as shown in Figure 4(b). For both equations, the
accuracy decreases when hM/R1 increases, probably because of
the turbulent fluctuations of the wave maximum for large Fo. It is
also noteworthy that the application of Equation (2), as conceived,
is allowable to predict hM for all the inlet structures with the
Figure 2. Supercritical flow along the spiral for test G: (a) top view and (b) first occurrence of shock waves (Jain 1984; Yu and Lee 2009; Quick
standing shock wave adhering to outer spiral wall. 1990).
228 G. CRISPINO ET AL.

Figure 5. Measured rotational flow angles α along vertical shaft.

a limit height zUF beyond which the vertical flow velocity


variation is negligible, and the quasi-uniform flow is
expected to establish, as
 
ð1=nÞ6=5 Q 4=5
zUF ¼ 3   (3)
g πDs

Herein n is the Manning roughness coefficient for the drop


shaft pipe. This coefficient can increase by some 10% to
account for pipe aging (Gisonni and Hager 2012).
The measured rotational flow angles α (angle between
locally visible stream line and the vertical) from the top
(z/Ds = 0.0) to the toe (z/Ds = 23.5) of the shaft are shown in

Figure 4. (a) Normalized maximum wave height hM/ho versus bend number B,
and (b) comparison between measured (subscript mis) and predicted (subscript
calc, Equations (1) and (2)) values of relative wave maxima hM/R1.

Vertical shaft
Centrifugal (rotational) and gravity (translational) forces act
on the falling flow along the vertical shaft. The rotational
effect originates from the spiral inlet constraining the flow
to assume an angular velocity. Due to the centrifugal force,
the flow adheres to the shaft wall establishing a helicoidal
flow. Once the flow is distant from the spiral inlet, the
rotation reduces because of friction and the vertical transla-
tion driven by gravity tends to prevail. Even if a perfect axial
flow asymmetry is hardly to be achieved along the shaft as
the flow rotation is dissipated, quasi-uniform flow (subscript
UF) is assumed to establish, at a certain distance from the
shaft top (z = 0.00), as soon as an equilibrium between
driving (gravity) and retarding (wall roughness) forces is
reached. The rotational effect is then almost lost (Hager Figure 6. n-factor used for increasing the resistance coefficient λ against the
and Kellenberger 1987). Kellenberger (1988) provided capacity Froude number FC.
URBAN WATER JOURNAL 229

Figure 5. The value αs = 60° at z/Ds = 0.0 corresponds to the other tests, the helicoidal flow still rotated (α > 0°) at the shaft
bottom angle of the spiral inlet and relative to the vertical. At end. So far, few analytical approaches for α, or alternatively for
a fixed shaft elevation, α increases by decreasing Fo or, corre- the axial (i.e., vertical) and tangential (i.e., horizontal) velocity
spondingly, by increasing the incoming discharge Q (Table 1). components Vv and Vh along the vertical shaft, are available
For Fo = 9.23 (test A) and Fo = 6.59 (test B), flow rotation lost up to the Authors knowledge (Jeanpierre and Lachal 1966; Liu
(α = 0°) at about z/Ds = 9.5 and z/Ds = 17.5, respectively. These et al. 2018). Kellenberger (1988) estimated the shaft velocity
no-rotation elevations are slightly larger (+25%) than zUF given magnitude as
by Equation (3), proving that the centrifugal effect is more
significant than that predicted by Kellenberger (1988). For all ðV=VUF Þ2 ¼ tanhðz=zUF Þ (4)

Figure 7. Measured rotational flow angles α along the vertical shaft for Fo = : (a) 9.23, (b) 6.29, (c) 4.61, (d) 3.36, (e) 2.50, (f) 2.37, (g) 2.36, (h) 2.30.
230 G. CRISPINO ET AL.

Figure 8. Normalized (a) Vv/Vv,3 and (b) Vh/Vh,3 profiles along the vertical shaft.

Herein, an analytical approach for computing α, Vv and Vh


Figure 9. (a) Free-surface mixture-flow profiles in the dissipation chamber, (b) flow
is proposed. The following assumptions and the conse- pattern for Fo = 2.50 (test E), and (c) hmax/Hch against FC.
quent results refer to the present VDS geometry, with
a spiral inlet at the shaft head. However, the present
The flow thickness e3 is instead obtained by applying the
procedure may be easily applied to VDSs controlled by
continuity equation:
other inlet devices (Yu and Lee 2009; Liu et al. 2018)
conveying swirling flows in the shaft. The flow entering e3 ¼ Q=ðπDs V3  cosαs Þ (5)
the vertical shaft is assumed to satisfy the free vortex
theory, according to which the angular momentum rV is Given two consecutive horizontal cross-sections of elevations
constant, with r as the rotational flow radius along the z and z + Δz, the momentum equation provides the velocity
spiral inlet. Given the approach flow velocity Vo and the components Vv,z+Δz and Vh,z+Δz (Jeanpierre and Lachal 1966) as
spiral end bottom slope αs, the resultant velocity V3 for r sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

  
= R3 (entrance in the vertical shaft) is thus known and, nλV z 2 V v;z 2
consequently, the axial (vertical) and tangential (horizontal) V v;zþΔz ¼ 1  cosαz   Δzþ  2g (6)
8ez g 2g
velocity components are V3v = V3∙sinαs and V3h = V3∙cosαs.
URBAN WATER JOURNAL 231

for z/Ds < 10.0 the flow rotation diminished more rapidly in
the physical model than predicted, probably because the
centrifugal force affects the flow regime more relevantly as
soon as the flow leaves the spiral inlet. A better prediction
of the observed data might be achieved by adopting an n-
factor decreasing from the head to the bottom of the shaft.
Figure 8 shows normalized Vv/Vv,3 and Vh/Vh,3 profiles
along the shaft. If a shaft elevation L/Ds is fixed, then the
increase of Fo, or equivalently the decrease of Q, leads to
a reduction of both the axial (vertical) and tangential (hor-
izontal) velocity components. The axial velocity component
increases under gravity along the shaft (Figure 8(a)) until
attaining a constant value (quasi-uniform flow), whereas the
tangential component gradually decreases due to friction
(Figure 8(b)).

Dissipation chamber
The hydraulic performance of the dissipation chamber is a key
aspect for the safety of a VDS. Few data are presently avail-
Figure 10. Average invert pressures p/Hu (from piezometers).
able, suggesting a standard box-shaped chamber of
Kellenberger (1988; see also Gisonni and Hager 2012).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
   Particularly, the invert pressure distribution remains unknown.
1 nλV z 2 V h;z 2 However:
V h;zþΔz ¼   cosαz   Δzþ  2g
cosαz 8ez g 2g
(7) ● maximum pressures due to the impact of the falling flow
(with sediments) are responsible for structural damage to
where αz, Vz, Vv,z and Vh,z are, respectively, the rotational the invert, and
flow angle, the velocity and the axial and tangential velocity ● pressure fluctuations create unstable outflow conditions,
components, at the horizontal cross-section with elevation with air entrainment and a risk to choke for the tailwater
z, and λ is the shaft resistance coefficient. Equations (6) and tunnel.
(7) allow, thus, to derive Vz and Vh profiles along the vertical
shaft by knowing their boundary values at the shaft In the present experimental campaign, the mixture flow
entrance. The rotation angle (relative to the vertical) at surface was visually derived. Then, the piezometers pro-
z is then obtained as vided average pressure values at x/Lch = 0.01 (T1), 0.12
  (T2), 0.24 (T3), 0.48 (T4), 0.72 (T5) and 0.96 (T6). In
αz ¼ arctan Vh;z =Vv;z (8) a second test campaign, the transmitters were mounted at
the same locations x/Lch for tests A to E (Table 1), recording
For each test, λ was computed by the approach of Zigrang
pressure time-series, and thus giving useful information
and Sylvester (1985) approach for the transition regime, in
about the pressure oscillations.
accordance with the prototype Reynolds number R = 4eV/ν,
The dissipation chamber showed an adequate perfor-
with ν as the water viscosity, ranging between about
mance in relation to the operation requirements described
6.21∙105 and 9.23∙106, and with the typical roughness
by Gisonni and Hager (2012). Under the tested discharges,
value for concrete pipes (Gisonni and Hager 2012). The
a breakdown of the air circulation was not observed. Time-
latter was adopted because the Cossonay VDS will be prob-
averaged mixture flow free-surface levels are shown in
ably realized by using the concrete material. In Equations (6)
Figure 9(a). The outlet (0.7 < x/Lch < 1.0) of the chamber
and (7), λ is multiplied by an n-factor to account for the
was submerged by mixture-flow for the ‘design’ and ‘over-
friction increase due to the centrifugal force (Akan 2006).
load’ (Table 1) tests as Fo ≤ 2.50 (Figure 9(b)), whereas this
This factor was assumed to be constant along the shaft and
free-surface was preserved for all other tests. The maximum
it was calibrated by minimizing the relative error between
mixture flow free-surface level hmax relative to the chamber
the measured and predicted rotation angles and prelimina-
height Hch is plotted against FC in Figure 9(c). If FC < 0.46,
rily defined as n = 10∙FC (Figure 6).
then the outlet of the chamber was not submerged and,
The rotation angle distribution as well as the axial and
according to the chamber observations, hmax/Hch can be
tangential velocity component profiles are respectively
estimated as:
shown in Figures 7 and 8, together with the measured
values. The angles from Equation (8) fit satisfactorily the hmax =Hch ¼ 1:50  ðFC Þ0:60 (9)
measurements for tests with large Froude numbers, i.e. Fo
≥ 3.36 (Figure 7(a–d)). A discrepancy is, instead, observed Contrarily, the outlet of the chamber was submerged
for overload tests with Fo ≤ 2.50 (Figure 7(e–h)). In general, (hmax/Hch = 1.0) for FC ≥ 0.46 (overload tests from E to
232 G. CRISPINO ET AL.

Figure 11. Boxplots of the dynamic invert pressures p/Hu (from transmitters), for Fo = (a) 9.23, (b) 6.29, (c) 4.61, (d) 3.36, (e) 2.50.

H, see Table 1). Given the design parameters Q and Ds and if a continuously decreasing pressure profile is assumed, then
if an acceptable value of hmax is assumed, then Equation the following relations may be utilized to predict the pressure
(9) may be adopted to preliminarily design the dissipation profiles in the dissipation chamber for 3.36 ≤ Fo ≤ 9.23 (tests
chamber height independently on the geometry compo- from A to D), with:
nents upstream of the vertical shaft.
Figure 10 shows the normalized average invert pressures p=Hu ¼ 0:07  ðx=Lch Þ0:20 (10a)
p/Hu, as registered by the piezometers. Here, Hu = zu + (Vu2/2g)
and for 2.50 ≤ Fo ≤ 2.30 (tests from E to H):
is the energy head of the falling flow, where Vu = (Vv,u2 + Vh,u
2 0.5
) was derived by Equations (6) and (7) and zu coincides with p=Hu ¼ 0:08  ðx=Lch Þ0:25 (10b)
the chamber height Hch. Pressures decrease as the degree of
liberty augments (when leaving x/Lch = 0.00), with maximum Both Equations (10a) and (10b) are valid for x/Lch ≥ 0.01.
at x/Lch = 0.01 at the shaft flow impact (still guided by the For larger Fo (tests from A to D), values of around p/Hu = 0.18
vertical chamber wall). Note that the flow does not rotate for are reached, indicating that only 18% of the incoming energy
smaller discharges (Figure 5), so that its impact is not located head was detected at the invert. For moderate Fo (tests from
at the chamber backwall (x/Lch = 0.01) but slightly shifted E to H), up to 32% of the incoming energy is transformed into
downstream between x/Lch = 0.01 and x/Lch = 0.24. However, pressure head on the invert. Downstream of the impact, values of
URBAN WATER JOURNAL 233

(i.e. Fo ≥ 6.29), whereas only one shock wave was detected for
moderate Froude numbers. The data analysis allowed for
proving that the maximum shock wave height is a function
of the bend number B. The resulting relation can be generally
employed for all inlet structures governed by deflecting flows
generating shock waves.
As regards to the vertical shaft, the measurement of the
rotational flow angle of the falling water indicated that the
rotational effect completely dissipated for Fo ≥ 6.29. For Fo ≤
4.61, the flow at the shaft toe still rotated, not showing a so-
called quasi-uniform flow condition. Based on the momentum
equation, an analytical description of the rotation angle and
velocity field along the shaft is presented. The latter approach
is applicable to vertical shafts controlled by whatever inlet
devices issuing a swirling flow. A future research to test the
reliability of the momentum equation approach will be desir-
able especially for overload conditions (e.g. Fo < 2.50).
The water level measurements in the dissipation chamber
highlighted that the chamber outlet was submerged for about
FC > 0.50. As regards to the pressure regime, the bottom of the
Figure 12. Dynamic minimum and maximum invert pressures p/Hu (from chamber just below the shaft outflow is a crucial region which
transmitters). should be conveniently protected against local damages due to
the flow impact. Here, the maximum pressure indicatively ranged
between 0.18 and 0.60 times the energy head Hu. Based on the
average pressure are equal to around 0.10∙Hu. This implies that
experimental observations, general relations for preliminary esti-
about 90% of the energy head at the shaft toe was dissipated,
mation of the chamber height and the pressure profile in the
mostly due to the jet impact on the invert.
chamber are proposed. These equations are theoretically valid
Boxplots of the dynamic pressures (recorded by the trans-
for standard types of vortex drop shafts with a dissipation
mitters) are shown in Figure 11. They indicate that the normal-
chamber.
ized pressure p/Hu measured at positions T1, T2 and T3 (shaft
outflow region) strongly fluctuated with a standard deviation
σ roughly equal to 0.15. Vice versa, there is hardly any fluctua- Acknowledgements
tion in the pressure values for positions T4, T5 and T6 (down-
stream portion of the chamber). Figure 12 shows the relative The support of the municipality of Cossonay and of the Laboratory of Hydraulic
Structures (LCH) of École Polytechnique Fédérale de Lausanne (EPFL), where the
pressure profiles p/Hu in the chamber from the transmitters.
experiments were conducted, is sincerely appreciated. In addition, Gaetano
Their measurements confirm that the maximum pressure is Crispino acknowledges the funding program “VALERE” (2018) by Università
located at x/Lch = 0.01 for 2.50 < Fo < 4.61 (e. g. tests C, D and della Campania Luigi Vanvitelli, which supported his research.
E), with p/Hu ranging between 0.30∙and 0.55∙Hu. Oppositely,
a relative (Fo = 9.23, e. g. test A) and an absolute (Fo = 6.29,
e. g. test B) maximum was measured at x/Lch = 0.24 when the Disclosure statement
flow leaving the shaft dissipated the rotational effect and it No potential conflict of interest was reported by the authors.
probably achieved the quasi-uniform flow. Overall, some 30%
to 60% of the inflowing head manifest as instantaneous max-
imum pressures at the chamber invert. However, more inter- ORCID
esting are negative pressures at the flow impact region below
G. Crispino http://orcid.org/0000-0002-3889-1115
the shaft. Values of −0.20∙Hu were recorded, pointing at
C. Gisonni http://orcid.org/0000-0002-9220-2149
a strong dynamic load of the invert with local negative pres-
sures that may cause the structural collapse also due to uplift
forces. References
Akan, A. O. 2006. Open Channel Hydraulics. Burlington, MA: Elsevier.
Camino, G. A., D. Z. Zhu, and N. Rajaratnam. 2014. “Flow Observations in
Conclusions Tall Plunging Flow Dropshafts.” Journal of Hydraulic Engineering 141 (1):
06014020. doi:10.1061/(ASCE)HY.1943-7900.0000939,06014020.
The results of an experimental vortex drop shaft investigation Crispino, G., D. Dorthe, T. Fuchsmann, C. Gisonni, and M. Pfister. 2016. “Junction
are herein presented to fill the existing gaps concerning the Chamber at Vortex Drop Shaft: Case Study of Cossonay.” In Hydraulic
hydraulic operation of this drop manhole. Eight tests, all with Structures and Water System Management, edited by B. M. Crookston and
a supercritical approach flow, were performed and analysed. B. P. Tullis, 437–446. Logan, UT: Utah State University, DigitalCommons.
De Martino, F., C. Gisonni, and W. H. Hager. 2002. “Drop in Combined Sewer
The flow regime of the supercritical spiral inlet is governed
Manhole for Supercritical Flow.” Journal of Irrigation and Drainage
by shock waves. Differently from Hager (1990), two shock Engineering 128 (6): 397–400. doi:10.1061/(ASCE)0733-9437(2002)
waves occurred for approach flows with large Froude numbers 128:6(397).
234 G. CRISPINO ET AL.

Del Giudice, G., C. Gisonni, and G. Rasulo. 2008. “Vortex Shaft Outlet.” In Liu, Z. P., X. L. Guo, Q. F. Xia, H. Fu, T. Wang, and X. L. Dong. 2018.
Advances in Water Resources and Hydraulic Engineering, edited by “Experimental and Numerical Investigation of Flow in a Newly
C. Zhang and H. Tang, 2053–2058. Beijing: Tsinghua University Press. Developed Vortex Drop Shaft Spillway.” Journal of Hydraulic
Del Giudice, G., C. Gisonni, and G. Rasulo. 2010. “Design of a Scroll Vortex Engineering 144 (5): 04018014. doi:10.1061/(ASCE)HY.1943-
Inlet for Supercritical Approach Flow.” Journal of Hydraulic Engineering 7900.0001444.
136 (10): 837–841. doi:10.1061/(ASCE)HY.1943-7900.0000249. Motzet, K. M., and F. Valentin. 2002. “Efficiency of a Vortex Chamber with
Drioli, C. 1969. “Installazioni Con Pozzo Di Scarico a Vortice [installations Horizontal Bottom under Supercritical Flow.” In Proceedings of IX
with a Vortex Drop].” L’Energia Elettrica 24 (10): 447–452. International Conference on Urban Drainage (9ICUD), 1–11. Portland,
Gisonni, C., and W. H. Hager. 2012. Idraulica Dei Sistemi Fognari: Dalla Oregon: American Society of Civil Engineers (ASCE).
Teoria Alla Pratica [wastewater Hydraulics: From Theory to Practice]. Pfister, M., and H. Chanson. 2014. “Two-phase Air-water Flows: Scale
Milan: Springer-Verlag. Effects in Physical Modeling.” Journal of Hydrodynamics 26 (2):
Hager, W. H. 1990. “Vortex Drop Inlet for Supercritical Approaching Flow.” 291–298. doi:10.1016/S1001-6058(14)60032-9.
Journal of Hydraulic Engineering 116 (8): 1048–1054. doi:10.1061/(ASCE) Pfister, M., G. Crispino, T. Fuchsmann, J. M. Ribi, and C. Gisonni. 2018.
0733-9429(1990)116:8(1048). “Multiple Inflow Branches at Supercritical-type Vortex Drop Shaft.”
Hager, W. H. 2010. Wastewater Hydraulics: From Theory to Practice. 2nd ed. Journal of Hydraulic Engineering 144 (11): 05018008. doi:10.1061/
Berlin: Springer. (ASCE)HY.1943-7900.0001530.
Hager, W. H., and M. H. Kellenberger. 1987. “Die Dimensionierung Des Quick, M. C. 1990. “Analysis of Spiral Vortex and Vertical Slot Vortex Drop
Wirbelfallschachtes [The Design of the Vortex Drop].” Gas- und Shafts.” Journal of Hydraulic Engineering 116 (3): 309–325. doi:10.1061/
Wasserfach (gwf). Wasser, Abwasser 128 (11): 582–590. (ASCE)0733-9429(1990)116:3(309).
Jain, S. C. 1984. “Tangential Vortex-inlet.” Journal of Hydraulic Engineering Reinauer, R., and W. H. Hager. 1997. “Supercritical Bend Flow.” Journal of
110 (12): 1693–1699. doi:10.1061/(ASCE)0733-9429(1984)110:12(1693). Hydraulic Engineering 123 (3): 208–218. doi:10.1061/(ASCE)0733-9429-
Jain, S. C. 1987. “Free-surface Swirling Flows in Vertical Dropshaft.” Journal (1997)123:3(208).
of Hydraulic Engineering 113 (10): 1277–1289. doi:10.1061/(ASCE)0733- Vischer, D. L., and W. H. Hager. 1995. “Vortex Drops.” In Energy Dissipators:
9429(1987)113:10(1277). Hydraulic Structures Design Manual, edited by CRC Press, 167–181. Vol. 9
Jain, S. C., and J. F. Kennedy. 1983. Vortex-flow Drop Structures for the (9). Rotterdam, The Netherlands: Balkema.
Milwaukee Metropolitan Sewerage District Inline Storage System. IIHR Rep. Yu, D., and J. H. W. Lee. 2009. “Hydraulics of Tangential Vortex Intake for
264. Iowa City, IA: Iowa Institute of Hydraulic Research, University of Iowa. Urban Drainage.” Journal of Hydraulic Engineering 135 (3): 164–174.
Jeanpierre, D., and A. Lachal. 1966. “Dissipation D’energie Dans Un Puits doi:10.1061/(ASCE)0733-9429(2009)135:3(164).
a Vortex [energy Dissipation across a Vortex Drop Shaft].” La Houille Zhao, C. H., D. Z. Zhu, S. K. Sun, and Z. P. Liu. 2006. “Experimental Study of
Blanche 7: 823–832. doi:10.1051/lhb/1966052. Flow in a Vortex Drop Shaft.” Journal of Hydraulic Engineering 132 (1):
Kellenberger, M. 1988. “Wirbelfallschächte in der Kanalisationstechnik [Vortex 61–68. doi:10.1061/(ASCE)0733-9429(2006)132:1(61).
drops in sewers].” PhD diss., Eidgenössische Technische Hochschule Zurich. Zigrang, D. J., and N. D. Sylvester. 1985. “Discussion of “A Simple Explicit
Knauss, J. 1987. Swirling Flow Problems at Intakes: Hydraulic Structures Formula for the Estimation of Pipe Friction Factor”. By J.J.J. Chen.”
Design Manual. Rotterdam: Balkema. Proceedings of the Institution of Civil Engineers 79: 218–219.

You might also like