Hydraulic Transients in Pressurized Pipe
Hydraulic Transients in Pressurized Pipe
IHE-Delft
Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Definitions and basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1 Time scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Classification of flow regimes . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3 Water-hammer wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3 Rigid column theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1 Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2 Derivation of equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4 Water-hammer theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1 Derivation of equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2 Simplified equations and numerical approaches . . . . . . . . . . . . . . . . 14
5 The method of characteristics (MOC) . . . . . . . . . . . . . . . . . . . . . 15
5.1 Characteristic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5.2 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6 Surge protection measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
6.1 Surge tank . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
6.2 Hydropneumatic vessel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.3 Valves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
6.4 By-passes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1
2
1 Introduction
A transient flow is the intermediate-stage flow, when the flow conditions are changed, from
one steady-state condition to another steady-state (Chaudhry, 2014). Valve manoeuvres,
starting or stopping of pumps, load acceptance or rejection of turbines, and a number
of other operations and processes occurring in pipe systems may induce fluid transients
which analysis is not possible by the steady or quasi-steady state approaches. When fluid
compressibility and pipe-wall elasticity become relevant, more advanced tools are required
for accurately describing unsteady flows in pipe conduits.
The present chapter aims at introducing the fundamentals of the classic water-hammer
theory. First basic definitions are described and a classification of flow regimes is provided.
Then, generalized forms of mass and momentum conservation equations are derived by
means of the Reynolds transport theorem, a simplified version is reached and expressions
for the wave speed are presented. After a brief explanation of the different numerical
approaches, the well-known Method of Characteristics (MOC) is explained and used to
solve the water-hammer equations. A section is devoted for surge protection methods,
which are explained by means of practical examples. Finally, a number of exercises are
suggested with the goal to better understanding the presented concepts.
• Steady flow: velocity is a constant value during all the considered period (i.e.,
V (t) = k). The representative time scale is as well t = ∞.
• Quasi-steady flow: velocity varies slowly enough to assume the conditions are
identical to those of a steady flow with the same instantaneous mean velocity. Rep-
resentative time scale: 2D/(f V ) t < ∞.
• Unsteady flow: two main approaches can be applied considering fluid inertia,
compressibility, and pipe-wall elasticity.
along the pipe can be assumed uniform (as long as the pipe cross-section is
constant). Representative time scale: t ∼ 2D/(f V ).
Wave speed
Wave speed, or wave celerity, is the velocity in which water-hammer waves propagate. This
propagation velocity depends on the bulk modulus of compressibility, the elastic properties
of the pipe-wall and the external constraints of the conduit. The general expression for
the water-hammer wave velocity is presented in Halliwell’s equation:
s
K
a= (1)
ρ[1 + (K/E)ψ]
in which ψ is a dimensionless parameter that depends upon the elastic properties of the
conduit. Following are presented the expressions for thin-walled conduits (e < D/10)
• Rigid conduits: ψ = 0
D
• Conduit anchored against longitudinal movement: ψ = e (1 − ν 2)
D
• Conduit anchored against longitudinal movement at the upper end: ψ = e (1−0.5ν)
D
• Conduit with frequent expansion joints: ψ = e
Wave period
The wave period depends on the celerity and the travelling distance of the propagating
water-hammer wave. For instance, in reservoir-pipe-valve systems the wave have to travel
3 Rigid column theory 6
two times back and forth the pipe to close a cycle (v.s. Fig. 1), while in valve-pipe-valve
systems it is the half. Hence, the wave period in a reservoir-pipe-valve system is:
4L
T = (2)
a
Wave amplitude
First presented in shock waves theory and later adapted for pressurized conduits, the
well-known Joukowsky equation relates pressure to velocity changes:
∆p = ρa∆V (3)
3.1 Assumptions
As mentioned in the previous section, rigid water-column theory assumes that the pipe-
walls are rigid and the liquid is incompressible. The theory solves the final steady state
3 Rigid column theory 7
for a perturbation caused in a pipe system, hence it is useful to determine the effects
of slow transient excitations (t ∼ 2D/(f V )). For faster transient excitations (e.g. fast
valve manoeuvres), elastic water-column theory (water-hammer theory) is required. For
an extended explanation of the rigid water-column theory see Parmakian (1963).
dV
ρgA (H0 + Ha ) − ρgA (H0 − Z1 ) − ρgAL sin α = −ρAL (5)
dt
Since L sin α = Z1 , rearranging Eq. 5 the momentum conservation equation of rigid
water-column theory is obtained (Eq. 6)
L dV
Ha = − (6)
g dt
4 Water-hammer theory
And the total amount of the extensive property B in the control volume is:
Z
Bcv = βρd∀ (14)
cv
The first term on the right hand side of Eq. 17 represents the time rate of change of
property B in the control volume, which taking into account Eq. 14 can be written as:
where A = cross-sectional area of the conduit V = flow velocity and W = velocity of the
control surface (see fig. 5).
Substituting Eqs. 18 and 19 into Eq. 17 the Reynolds transport theorem takes the
following form:
Z
dBsys d
= βρd∀ + [βρA(V2 − W2 )] − [βρA(V1 − W1 )] (20)
dt dt cv
4 Water-hammer theory 10
∆m
β = lim =1 (21)
∆m→0 ∆m
d x2
Z
ρAdx + ρ2 A2 (V2 − W2 ) − ρ1 A1 (V1 − W1 ) = 0 (22)
dt x1
Noting that dx2 /dt = W2 and dx1 /dt = W1 , the previous equation simplify to
Z x2
∂
(ρA)dx + (ρAV )2 − (ρAV )1 = 0 (24)
x1 ∂t
1 dρ 1 dA ∂V
+ + =0 (28)
ρ dt A dt ∂x
The previous Eq. 28 is already the continuity equation, though, the derivatives of ρ
and A must be expressed in terms of the variables of interest, which are V and P .
For this purpose, first we define the bulk modulus of elasticity, K, of a fluid by
dP
K= (29)
dρ/ρ
so the derivative of ρ in Eq. 28 can be written as
dρ ρ dP
= (30)
dt K dt
On the other side, for a circular conduit having radius R,
dA dR
= 2πR (31)
dt dt
In terms of circumferential strain
dA 1 dR
= 2πR2 (32)
dt R dt
or
1 dA d
=2 (33)
A dt dt
Assuming the pipe material behaviour is linearly elastic and that there are expansion
joints throughout the pipe, the circumferential strain in terms of circumferential stress
σ is
σ
= (34)
E
And the circumferential stress in a thin-walled conduit for inner pressure P load is
Pr
σ= (35)
e
where e = thickness of the conduit. By taking time derivative
dσ P dr r dP
= + (36)
dt e dt e dt
Based on Eq. 35 the previous expression can be written as
d P r d r dP
E = + (37)
dt e dt e dt
which can be simplified to
r dP
d
= e dtP r (38)
dt E− e
4 Water-hammer theory 12
Substituting Eq. 42 into Eq. 41 and in terms of partial derivatives the continuity
equation takes the following form:
∂P ∂P ∂V
+V + ρa2 =0 (43)
∂t ∂x ∂x
(ρAV 2 )2 − (ρAV 2 )1
P
∂ F
(ρAV ) + = (49)
∂t ∆x ∆x
Forces acting on the system in the pipe axial direction (Fig. 6)
where θ = is the complementary angle between gravity direction and pipe axis, and
τ0 = shear stress between the conduit and the fluid.
The balance of forces considering downstream direction as positive is
X 1
F = P1 A1 − P2 A2 − (P1 + P2 )(A1 − A2 ) − ρgA(x2 − x1 ) sin θ − τ0 πD(x2 − x1 ) =
2
1
= (P1 − P2 )(A1 + A2 ) − ρgA(x2 − x1 ) sin θ − τ0 πD(x2 − x1 )
2
(50)
In the present derivation steady friction head losses are assumed and Darcy-Weisbach
friction formula is used
1
τ0 = ρf V |V | (53)
8
where f = Darcy-Weisbach friction factor.
Substituting Eq. 53 into Eq. 52 and expanding terms
∂ ∂V ∂ ∂V ∂P ρAf V |V |
V (ρA) + ρA +V (ρAV ) + ρAV +A + ρgA sin θ + = 0 (54)
∂t ∂t ∂x ∂x ∂x 2D
Rearranging gives
∂ ∂ ∂v ∂V ∂P ρAf V |V |
V (ρA) + (ρAV ) + ρA + ρAV +A + ρgA sin θ + = 0 (55)
∂t ∂x ∂t ∂x ∂x 2D
According to the continuity equation Eq.26 the sum of the terms in the squared brack-
ets is zero. Hence dropping the terms and dividing by ρA, the momentum conservation
equation is reached
∂V ∂V 1 ∂P f V |V |
+V + + g sin θ + =0 (56)
∂t ∂x ρ ∂x 2D
dQ gA dH
± + RQ|Q| = 0 (65)
dt a dt
Note that the compatibility equations, Eqs. 65, are only valid for dx dt = ±a. Hence
delimiting the solution over straight lines with slope ±1/a one independent variable can
be cancelled out and the partial differential equations are converted to ordinary differential
equations. The lines over the solution is possible are called characteristic lines (see Fig.7)
and physically they represent the path over the xt ˆ plane traversed by a disturbance.
5 The method of characteristics (MOC) 16
ˆ plane.
Fig. 7: Characteristic lines in the xt
Based on Fig.7 Eqs. 65 can be solved by multiplying the left hand side by dt and
integrating Z P
gA P
Z Z P
dQ + dH + R Q|Q|dt = 0 (66)
A a A A
and Z P Z P Z P
gA
dQ − dH + R Q|Q|dt = 0 (67)
B a B B
Although the first two terms in the previous equations are easy to evaluate, an ap-
proximation must be carried out for the friction losses term. Following, a first order
approximation (Eq. 68) and a second order approximation (Eq. 69) are presented.
Z P
R Q|Q|dt ≈ RQA |QA |∆t (68)
A
Z P
R Q|Q|dt ≈ R∆t|QA |QP (69)
A
For the sake of simplicity in the present analysis a first-order approximation is applied
allowing an explicit solution of the system. The first-order approximation usually yields
satisfactory results for most engineering applications.
Finally the characteristic equations are reached
gA
QP − QA + (HP − HA ) + R∆tQA |QA | = 0 (70)
a
gA
QP − QB − (HP − HB ) + R∆tQB |QB | = 0 (71)
a
By combining the known variables together Eqs. 70 and 71 the positive and negative
characteristic equations are reached:
QP = Cp − Ca HP (72)
QP = Cn + Ca HP (73)
5 The method of characteristics (MOC) 17
being
and
gA
Ca = (76)
a
Solving explicitly the system of equations (Eqs. 72 and 73) HP and QP can be de-
termined in each time step. Though, notice that the characteristic equations are only
valid along the characteristic lines AP and BP respectively. The constants Cp and Cn
are known values, they represent the information transferred from time step to time step
through the positive and negative characteristics lines. And Ca depends upon the conduit
properties.
Constant-level upstream
In many situations the changes in the reservoir level are negligible. If the entrance losses
as well as the velocity head are negligible, then
in which Hres = height of the reservoir water surface. Eq. 73 for the upstream end
thus becomes
QP1,j = Cnj + Caj Hres (78)
QPn,j = 0 (79)
Cp j
HPn,j = (80)
Caj
5 The method of characteristics (MOC) 18
(Q0n,j τ )2
Q2Pn,j = HPn,j (83)
H0n,j
By means of the positive characteristic equation (Eq. 72) the previous equation becomes
in which Cv = (τ Q0n,j )2 /(Ca H0n,j ). Solving for QPn,j and neglecting the negative sign in
the radical term q
QPn,j = 0.5(−Cv + Cv2 + 4Cpi Cv ) (85)
And finally HPn,j is obtained applying the positive characteristic equation (Eq. 72). Notice
that τ is the parameter that defines the valve manoeuvre. τ = 1 corresponds to the initial
state valve opening and τ = 0 corresponds to a closed valve. Values of τ during the
manoeuvre transition may be specified by an algebraic expression or in a tabular form.
Branching junctions
Branching junctions are solved by applying mass and energy conservation principles at
the branching node. Considering Fig. 8, mass conservation equation can be expressed as
Fig. 9: Picture of a surge tank protecting a water transport system, Tortosa (Spain).
Depending on their design and site constraints they may take several forms:
• Simple tank: open shaft or standpipe connected to the pipeline with negligible losses
at its junction with the pipeline (v.i. Fig. 10-a).
• Orifice tank: simple tank with an orifice restricting the entrance to the surge tank
(v.i. Fig. 10-b).
• Differential tank: two composed surge tanks, a simple surge tank and an orifice surge
tank (v.i. Fig. 10-c).
• One-way tank: the water is allowed to flow only from the tank to the main pipe with
the goal to protect the system for pressure drops (v.i. Fig. 10-d).
• Closed tank: with its top closed and partially filled by compressed air (v.i. Fig. 10-e).
• Tank with galleries: with upper or lower galleries, providing larger cross-sectional
area and storage capacity (v.i. Fig. 10-f).
6 Surge protection measures 21
Where L i the pipe length, Ap the cross-sectional area of the pipe and As the cross-sectional
area of the surge tank.
Generally, the time-scale of a water-hammer event is much shorter than the time-scale
of the mass oscillation produced by surge tanks, hence the two problems use to be treated
separately. In Streeter & Wylie (1967) and Chaudhry (2014) methods for modelling and
sizing surge tanks are described.
6.3 Valves
Hydraulic transients can be controlled by means of auxiliary valving and flow control
valves. The aim is mainly to allow a rapid outflow for a certain upsurge pressure or an air
inflow for a downsurge pressure. Several types of valves may be used for transient control:
• Safety valves (Fig. 13): they open when the pressure inside the pipe exceeds the
pressure limit by a spring or weight-loaded mechanism. They close rapidly when the
pressure drops bellow the limit.
• Air-inlet valves (Fig. 15): they let air flow in the pipe avoiding cavitation and the
high pressures derived from column separation.
• Check valves (Fig. 16): used to prevent reverse flow through a pump and to prevent
inflow into a one-way surge tank from the pipeline.
6.4 By-passes
By-pass valving is a low-cost efficient technique for avoiding pressure surges (up and down)
in pumping stations. Several configurations are possible. Essentially the aim is to give
continuity to the water flow in case of pump failure. Figure 17 depicts a typical by-pass
set-up in a pumping station.
7 Exercises
Joukowsky pressure rise: Joukowsky equation gives the piezometric head rise resulting
from an instantaneous closure of the valve in a frictionless system for an initial steady flow
aV0
∆H = (93)
g
Notice that using MOC positive characteristic equation and forcing QP = 0, the previous
expression is easily obtained. Derive it and compute Joukowsky pressure according to
the celerities obtained in the previous exercise and considering an initial flow velocity
V0 = 0.3 m/s.
Assuming copper is the material of the reservoir-pipe-valve system, use the excel
spreadsheet from the lecture material and run three different simulations for a linear
closure of the valve considering the following closing times: Tf = 0, Tf = 2L/a and
Tf = 4L/a. The initial flow conditions before closing the valve are: piezometric head
in the reservoir Hres = 40 m, piezometric head in the valve Hv = 39 m and initial flow
7 Exercises 28
1. Compare the performance upstream the valve (N8) with: no protection devices, a
surge tank (D = 0.5m) and a hydropneumatic vessel (V = 0.785m3 ).
Discuss the different results obtained, what are the advantages of hydropneumatic vessels
vs. surge tanks?
4. Simulation-4: change the water surface elevation of one of the upstream reservoirs
to Hres = 3m.
Compare and discuss the pressure output from the four tests at the downstream section
of pipe-3.
2. Incident: simulate a valve trip of the upstream gate for a closing time of 20 s
and the downstream closing after 244 s. This is actually the incident described in
(Purcell, 2015). Why did the hydropower scheme failed?
3. Solution: design a solution to avoid the problems observed at the previous simula-
tion.
7 Exercises 33
References