Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Case Studies in Thermal Engineering 49 (2023) 103306

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


journal homepage: www.elsevier.com/locate/csite

Optimization of end-wall fence in turbine based on response


surface methodology and genetic algorithm
Xu Han *, Qiuliang Zhu, Jiandong Guan, Zhongwen Liu, Bochuan Yao, Zhonghe Han
Hebei Key Laboratory of Low Carbon and High-Efficiency Power Generation Technology, North China Electric Power University, Baoding, 071003,
Hebei, China

A R T I C L E I N F O A B S T R A C T

Handling Editor: Huihe Qiu Secondary flow loss accounts for a large proportion of the internal flow loss in turbine stages. The
Keywords:
use of end-wall fences can effectively reduce secondary flow loss. In this study, the White cascade
Turbine was taken as the research object, and the position of the end-wall fence was parameterized. Based
Wet steam on the response surface method, the mapping relationship between the fence position and the
End-wall fence isentropic expansion efficiency was obtained, and a surrogate model was constructed. Finally,
Response surface method single-objective and multiobjective optimizations were carried out using genetic algorithms to
Genetic algorithm obtain the optimal fence position parameters. The results showed that the existence of the fence
Shock wave can not only reduce secondary flow but also effectively reduce shock losses when the fence is
located at the rear of the passage. Therefore, the optimization effect of the blade position is best
when it is located at the end of the passage. When the flow deviates significantly from the design
condition, the end-wall fence can significantly reduce the low-speed region on the pressure side of
the fence. Therefore, the fence can play a greater role in low-load conditions with significant
deviations from the design condition. After optimizing the design conditions, when H is 128.6
mm, V is 66.93 mm, and A is 56.48◦ , the isentropic expansion efficiency is the highest, reaching
96.160%. After optimizing the multi-inlet angle at low load conditions, when θ is 56.07◦ , R is
129.7 mm, and A is 64.36◦ , the comprehensive optimization result is the best, and the isentropic
expansion efficiencies at inlet angles of 0◦ , 10◦ , and 45◦ are 96.098%, 96.050%, and 93.930%,
respectively. The research results can provide a reference for the design of turbine flow.

Nomenclature

e energy density, J/(kg⋅m3)


Gb produced by turbulent kinetic energy
g volume force of gravity, N
hfg latent heat of vaporization, J/kg
ht total enthalpy, J/kg
I nucleation rate, kg− 1⋅s− 1
Kb Boltzmann constant, 1.38 × 10− 23 J/K
Kn Knudsen number
k turbulent kinetic energy, J

* Corresponding author.
E-mail address: xuhan@ncepu.edu.cn (X. Han).

https://doi.org/10.1016/j.csite.2023.103306
Received 17 April 2023; Received in revised form 12 July 2023; Accepted 15 July 2023
Available online 16 July 2023
2214-157X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Mm molecular weight of water, kg


ṁ mass condensation rate, kg/(m3⋅s)
N number of droplets per unit volume
Pr Prandtl constant
p static pressure, Pa
qc condensation coefficient
R gas constant of water vapor
r droplet radius, m
r* critical radius, m
s droplet surface correction coefficient
T temperature, K
ΔT degree of subcooling, K
t time, s
u velocity, m/s
x cartesian coordinate, m
Y wetness

Greek
Ω mean rate of rotation tensor
γ specific heat capacity ratio of steam
ε dissipation rate of turbulent kinetic energy
ζ resistance coefficient
η Isentropic expansion efficiency
λ thermal conductivity, W/(m⋅K)
μ dynamic viscosity, N⋅s/m2
ν kinematic viscosity, m2/s
ρ fluid density, kg/m3
σ surface tension, N/m
τ stress, Pa
τr relaxation time, s
φ droplet growth correction coefficient
ψ shock intensity
ω angular velocity, rad/s

Subscripts
g vapor phase volume average parameter
i,j,k unit vector
l liquid phase volume average parameter
m two-phase average parameter
s saturation parameter
t turbulent flow parameter

Superscript
− average parameter
~ true value in each phase
→ vector symbol

1. Introduction
The efficiency of a steam turbine is fundamentally limited by the Carnot cycle [1], which can be effectively improved by increasing
the main and reheat steam parameters and decreasing the exhaust pressure [2]. However, the main and reheated steam temperatures
of the steam turbine are limited by the blade material [3]. Increasing the main steam pressure would lead to an increase in flow rate,
resulting in higher humidity in the last few stages. Reducing the exhaust parameters would increase the power consumption of the
vacuum pump, which is unfavorable for its economic efficiency [4]. Therefore, to further improve the efficiency of steam turbines,
scholars have focused on flow passage, reducing flow losses and improving operational efficiency [5–8].
The flow losses in the flow passage of the steam turbine mainly include losses due to blade profile, secondary flow, and steam
leakage [9], with the secondary flow losses accounting for approximately 30% of the total; hence, its energy loss cannot be ignored
[10]. As shown in Fig. 1, secondary flow is formed because there is a pressure gradient along the transverse direction of the passage.
The fluid located in the boundary layer near the end wall is unable to balance the transverse pressure difference due to its low kinetic
energy, causing the fluid in the boundary layer to flow from the pressure surface to the suction surface of the blade. Controlling the

2
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 1. Vortex structure in the turbine cascade [11].

secondary flow loss on the end wall is the key to improving the flow efficiency. Moon [11], Kawai [12], Kumar [10] and others have
achieved good results by increasing the height of the end-wall fence, finding that the effect is best when the fence height is 1/3 of the
inlet boundary layer height, with a total pressure loss reduction of 15%. By using nonaxisymmetric end-wall shaping to suppress
secondary flow, Poehler [13] optimized the 3D design of a 1.5-stage turbine and unexpectedly found that the nonaxisymmetric
end-wall shaping rearranged the secondary flow pattern, resulting in an increase in stator loss but a decrease in rotor secondary flow,
ultimately leading to an increase in stage efficiency.
The lateral flow in the cascade becomes more pronounced when deviating from the design operating conditions. In the context of
global low-carbon goals [14], new energy generation technologies are rapidly developing. To accommodate new energy generation,
thermal power units need to increase their deep peak-shaving capacity to ensure safe, stable, and economical operation of the power
system [15–17]. Under deep peak-shaving conditions, operating at low loads affects the safety and economy of the steam turbine.
Retaining only a small amount of steam entering the low-pressure cylinder greatly deviates the flow state from the design operating
condition, increasing flow instability [18] and making it prone to flutter, greatly reducing its efficiency. To suppress the transverse
flow deviation from the design operating conditions, the addition of end-wall fences can be implemented. However, optimizing the
location of the end-wall fences in the flow field is necessary to better reduce flow losses.
Optimization design first requires a clear understanding of the impact of structural and flow parameters on flow losses and then
designing corresponding optimization schemes to further reduce flow losses. Currently, widely used optimization algorithms include
genetic algorithms (GA), taboo search algorithms (TS), simulated annealing algorithms (SA), particle swarm optimization (PSO),
sequential quadratic programming algorithms (NLPQL), etc. [19–22]. Yan C [23] optimized a turbofan aircraft engine, considering
three complex coupled subsystems, six coupled rules, fifty-four design variables, and nine constraints, and used NLPQL to quickly find
the optimal solution in the fan subsystem and low-pressure turbine subsystem. Shojaeefard [24] optimized the inducer of a centrifugal
pump using NSGA-II to perform multiobjective optimization design on the angle of the inlet blade tip, angle of the outlet blade tip, and
ratio of the outlet hub radius to the inlet hub radius. Finally, the coefficients of the water head, hydraulic efficiency, and NPSHR
increased by 14.3%, 0.3%, and 30.2%, respectively. Peng K [25] used an adaptive simulated annealing algorithm to optimize the guide
vane regulator of an aircraft engine compressor, obtaining good static and dynamic characteristics. Different optimization algorithms
can achieve good optimization results, and this paper chooses a genetic algorithm, which has a wide range of applications and can
perform single-objective and multiobjective optimization.
For the optimization research of the end-wall fence, Cho J [26] and Chen YZ [27] optimized the end-wall fence, obtaining
maximum efficiency by changing the fence thickness, fence height, and fence width. Uehara R [28] optimized the shape, installation
position, and setting angle of a 3D fence to alleviate the interaction between the horseshoe vortex and the end-wall cross-flow. Our
research group [29] combined the end-wall blade and heating and dehumidification methods to design three different fence positions,
obtaining an optimal result. Compared with the average outlet wetness, total pressure loss, and entropy production of the optimized
cascade, they decreased by 43.4%, 2.0%, and 2.0%, respectively. However, the determination of the fence position in this study is
relatively arbitrary, and the best optimization result cannot be obtained. Moreover, the above studies did not design optimization for
low load operating conditions. Therefore, based on the above, this paper uses response surface methodology to obtain the mapping
relationship between each optimization parameter and the optimization objective and constructs a proxy model for the fence position
and optimization objective. The attack angle generated by deviation from the design conditions under different inflow angles is
considered to simulate the effect of the fence in low-load operating conditions. Finally, using genetic algorithms for single and
multiobjective optimization, the best fence position parameters are obtained. The research results can provide a reference for the
design of turbine flow.

2. Model establishment
2.1. Mathematical model
This paper employs the Euler-Euler multiphase flow model to describe the nonequilibrium spontaneous condensation process of

3
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

wet steam [30,31]. The vapor and liquid phases have their respective governing equations, and the interphase interaction is added to
the control equation group in the form of source terms [32,33].
Continuity equation:

⎪ ∂ρg ∂ ( )

⎨ ∂t + ∂xj ρg uj = − ṁ

(1)

⎪ ∂ρm Y ∂ ( )

⎩ + ρm Yulj = ṁ
∂t ∂xj
Conservation equation for momentum in the vapor phase:
( ) ( )
∂ ρg u i ∂ ρg ui uj ∂p ∂τij ρ Y
+ =− + + Δρg gi + m (uli − ui ) − ui ṁ (2)
∂t ∂xj ∂xi ∂xj τrl ς
Conservation equation for momentum in the liquid phase:
( )
∂(ρm Yuli ) ∂ ρm Yuli ulj ρ Y
+ = ρm Ygi + m (ui − uli ) + (ui − uli )ṁ (3)
∂t ∂xj τrl ς
[ ( )]
1.74
ζ = 1 + Kn 2.492 + 0.84 exp − (4)
Kn
The energy conservation equation is given by:
( ) ( ) ( )
∂( ) ∂ ( ) ∂ ∂T ∂ puj ∂ τkji uki
ρg e + ρg uj e = λ − + + ṁhfg − ṁht (5)
∂t ∂xj ∂xj ∂xj ∂xj ∂xj
This paper uses the nonisothermal correction model proposed by Kantrowitz [34] to calculate the nucleation rate.
[( ) ]
∂(ρε) ∂(ρεui ) ∂ μ ∂ε ε ε2
+ = μ+ t + G1ε (Gk + G3ε Gb ) − C2ε ρ + Sε (6)
∂t ∂xi ∂xj σ ε ∂xj k k

The expression of η is:


√̅̅̅̅̅̅̅̅̅ ( )
ρ R RTg hfg hfg 1
η = qc g − (7)
α 2π RTg RTg 2

Gyarmathy [35] proposed a suitable model for droplet growth based on Fick’s diffusion coefficient and Fourier’s thermal con­
ductivity coefficient:
dr λg 1 − r∗ /r
= · · ΔT (8)
dt ρl hfg r(1 + 3.18Kn)

This paper adopts the model proposed by Young [36] to describe the process of droplet growth, which has been modified for the
low-pressure region below 30 kPa and is suitable for the operating conditions studied in this paper.
dr λg 1 − r∗ /r
= · 1 Kn
· ΔT (9)
dt ρl hfg r 1+2βKn + 3.78(1 − ξ) · Prg

[ ]
RTs 2 − qc γ + 1 RTs
ξ= · ψ (P) − 0.5 − · · (10)
hfg 2qc 2(γ − 1) hfg
[ ( )]
P
ψ (P) = 3.25 1 − tanh − 2 (11)
104
Mass condensation rate [37,38]:

4πr∗3 dr
ṁ = J ρl + 4πr2 ρl N (12)
3 dt
This paper uses the model proposed by Marcus Mccallum and Roland Hunty [39] to calculate the surface tension of liquid droplets.
( )
σ = s 85.27 + 75.612Td − 256.889Td2 + 95.928Td3 × 10− 3 (13)

In this paper, the turbulence model adopts the k-ε model with swirl correction, and the kinetic energy equation of vapor phase tur­
bulence is:

4
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 2. Grid partitioning and model verification.

( ) ( ) ( ) ( )
∂ ρg k ∂ ρg k̃u ∂ ̃uj ∂ 1 ′′ ′′ ′′ ′ ∂ ( ′′ ′′ )
+ = − ρg u′′i u′′j − ρg ui uj uj + ρ u′′i − τ u +
∂t ∂xi ∂xi ∂xi 2 ∂xi ij j
(14)
′ ∂u′′ ∂u′′j 1 [ ( ′ ′′ ) ( ) ′ ]
p − τ′′ij − ρm ṁk + ρl ulj uj − 2k + ulj − ̃uj ρl u′′j
∂xi ∂xi τrl ς
The turbulent kinetic energy equation of the liquid phase is:
( ) ( ) ( )
∂(ρl kl ) ∂ ρg kl ̃uli ′ ′ ∂ ̃
ulj ∂ 1 ′ ′ ′ ′ ′
+ = − ρl uli ulj − ρl uli ulj ulj + gj ρl uli +
∂t ∂xi ∂xi ∂xi 2
( ) (15)
1 [ ( ′ ′′ ) ( ) ′ ′]
ρm ṁ + ρl ulj uj − 2kl − ulj − ̃uj ρl ulj
τrl ς
Since the interphase interaction has a certain influence on the turbulent dissipation, the ε equation is modified:
[ ]
∂(ρε) ∂(ρεui ) ∂ (μ + μt ) ∂ε ε2
+ = + ρC1 S − 1.9ρ √̅̅̅̅̅+
∂t ∂xi ∂xi σk ∂xi k + νε
(16)
ε
(1.728Gb + 1.44Gl )
k

where

5
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 3. Dimensions and parametrization of the end-wall fence.

[ ] ( )
Sk ( )1/2 1 ∂ui ∂uj
C1 = max 0.43, , S = 2Sij Sij , Sij = + ;
Sk + 5ε 2 ∂xj ∂xi
( ⃒ ⃒/⃒ ⃒ )
1 ⃒→ ⃒ ⃒→⃒
Ql = 2⃒⃒ V l ⃒⃒ ⃒ V ⃒k − 2k − 2ṁk.
τrl ς
The turbulent viscosity coefficient of the vapor phase is:

k2
μt = ρ C μ (17)
ε
For the k-ε model with swirl correction, Cμ is no longer a constant in the above formula, which is calculated by the following
formula:
1
Cμ = √̅̅̅ (18)
4+ 6cos φU ∗ kε

where
(√̅̅̅ ) √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
1 ̃ ij Ω̃ ij ;
φ = arccos 6 W , U ∗ = Sij Sij + Ω
3

Sij Sjk Ski ̃


W = √̅̅̅̅̅̅̅̅̅ , Ω ij = Ωij − 2εijk ωk , Ωij = Ωij − εijk ωk .
Sij Sij

2.2. Cascade model


This paper focuses on controlling and optimizing the secondary flow loss of the White cascade, which is the stator of the last stage of
a 660 MW steam turbine. For parameters of the cascade, blade pitch is 87.59 mm, blade chord is 137.51 mm, stagger angle is 45.32◦ ,

6
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Table 1
Comparison of three simulation cases and ORI.

A (◦ ) H (mm) V (mm) η
Case 1 20 80 115 95.818%
Case 2 50 115 80 95.894%
Case 3 65 130 45 95.960%
ORI – – – 95.402%

inlet flow angle is 0.0◦ , inlet and leading edge are 30.76 mm apart. The cascade has an outlet design Mach number of 1.2 and an outlet
incidence angle of approximately 71◦ . It was found in experiments that the cascade suffers from high losses due to factors such as
condensation of wet steam, and thus, a deeper study was conducted [40].
To improve the computational accuracy, a hexahedral structured grid was used for the flow channel, with grid refinement applied
to the wall boundary layer, leading edge, and trailing edge, as shown in Fig. 2(a). Furthermore, the computational results were
validated against experiments [40] and grid independence, where the inlet total temperature was 354 K, inlet total pressure was 40.9
kPa, and outlet average static pressure was 15.3 kPa. The pressure distribution on the blade surface was calculated for grid numbers of
1.25 × 105, 2.38 × 105, 3.56 × 105, and 4.58 × 105. The comparison between the numerical simulation and experimental results is
shown in Fig. 2(b), which indicates that the number of grids has a significant effect on the pressure distribution on the suction surface
of the blade, where the influence of shock waves results in insufficient pressure rise for models with a small number of grids. When the
grid number exceeds 3.56 × 105, the pressure change becomes relatively small. Considering the balance between computational speed
and accuracy, the model with 3.56 × 105 grids is suitable for subsequent studies. In addition, the computational results were found to
be in good agreement with the experimental results, except for errors caused by interference from shock waves and wakes at the
trailing edge.

3. End-wall fence design


For the end-wall fence design, as shown in Fig. 3(a), the chord length of the fence is 41 mm, the thickness of the leading edge is 2
mm, and the thickness of the trailing edge is 1 mm. In the axial direction of the fence, the fence thickness increases linearly at first and
reaches the maximum thickness of 3 mm at 33.3% of the axial chord length, and then the fence thickness decreases linearly to 1 mm.
The trailing edge of the fence is a plane, and to reduce upstream losses, the leading edge is a circular arc with a radius of 2.2 mm. The
shape of the fence is fixed in this paper, and its position is parameterized by the horizontal distance (H) and vertical distance (V) of the
fence relative to the origin, as well as the angle (A) between the fence and the x-axis, as shown in Fig. 3(b).
According to the different positions of the end-wall fence, three operating conditions were simulated, representing the flow con­
ditions when the end-wall fence is located at the front, middle, and rear of the passage. The operating condition without an end-wall
fence is referred to as ORI. The position parameters of the end-wall fence and its isentropic expansion efficiency are shown in Table 1.
The end-wall fence is helpful in improving the overall efficiency. The reason why the overall improvement is not significant is that the
efficiency of the ORI is already high, reaching 95.402%. The condition with the greatest efficiency improvement among the three
operating conditions reached 95.960%, which is not easy to achieve and is quite significant. Fig. 4(a) shows that the end-wall fence can
promote streamline flow along the passage at the end, which plays a crucial role in hindering the formation of secondary flow.
Fig. 4(b) shows the Mach number distribution cloud maps for the three operating conditions. For ORI, the steam undergoes pressure
reduction expansion within the blade passage, but due to the influence of shock waves, the Mach number gradually increases and then
rapidly decreases, and after the steam continues to expand, it encounters another shock wave, and the Mach number decreases rapidly
again. After adding the end-wall fence, for Case 1, the fluid at the end-wall fence location did not reach supersonic, nor did it condense,
so no aerodynamic shock waves and condensation shock waves were generated. The efficiency improvement mainly comes from
suppressing the secondary flow. For Case 2 and Case 3, the end-wall fence strengthened the degree of steam expansion and acceler­
ation. Steam on the pressure side of the end-wall fence in Case 2 also reached supersonic. The wake of the end-wall fence in Case 2
interrupted the original shock wave, while in Case 3, the end-wall fence itself blocked the shock wave.
To better capture the structure of shock waves, a parameter, namely, the shock wave intensity ψ , was defined.
u i Pi
ψ= (19)
aP

In the equation, a is the local speed of sound, m/s; u is the velocity, m/s; P is the pressure, Pa; and i is the unit vector. Therefore,
dimensional analysis indicates that the shock intensity, ψ , is a scalar quantity. The numerical value of the shock intensity in the region
where the fluid is compressed is positive (ψ > 0), and if a shock wave is formed in this region, the numerical value of the shock intensity
is greater than 1 (ψ > 1). The numerical value of the shock intensity in the region where the fluid is expanded is negative (ψ < 0), and if
an expansion wave or beam wave is formed in this region, the numerical value of the shock intensity is less than − 1 (ψ < -1). The shock
intensity essentially represents the projection of the Mach number of the incoming flow in the direction of the pressure gradient in the
flow field, the degree of change of the pressure gradient along the velocity direction, which is consistent with the definition of the
normal Mach number, Ma, orthogonal to the shock wave.
Using the shock intensity to create a shadowgraph image, as shown in Fig. 4(c), the white area in the image represents shock waves
with intensity greater than 1. It can be clearly seen that a "λ"-shaped shock wave appears at the trailing edge of the blade. The dis­
tribution of the shock waves in case 1 is almost the same as that in ORI. In case 2, the original shock intensity has decreased, but a new

7
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 4. Comparison of genetic algorithm optimization results.

shock wave has appeared on the pressure side of the end-wall fence. As for case 3, it can be clearly seen that the presence of the end-
wall fence at the trailing edge has interrupted the original shock wave. Although the end-wall fence itself also produces a shock wave,
this is unavoidable. From the perspective of overall efficiency improvement, the end-wall fence near the trailing edge can not only
suppress the secondary flow but also greatly reduce shock wave losses. Therefore, case 3 has the highest efficiency among the three
cases.

8
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 5. Genetic algorithm optimization design.

4. Optimal design
To fully leverage the improvement in flow efficiency brought by the end-wall fence, the optimization of the fence position is carried
out in the following process shown in Fig. 5(a). Prior to optimization, the position of the end-wall fence is parameterized. The input
parameters for the optimization design are the horizontal distance (H), vertical distance (V), and angle (A) of the end-wall fence
relative to the coordinate origin. The output parameter is the isentropic expansion efficiency (η), and the constraint condition is
defined by equation (20), which requires the end-wall fence to be confined within the channel, as shown in Fig. 3(b). The optimization
process is divided into two parts: response surface construction and genetic algorithm optimization. First, sampling is carried out using
Latin hypercube sampling, which uses a layered, sampled, and shuffled method to ensure the comprehensiveness and uniformity of the
sample results with fewer sample points. Then, the sample points are calculated, and the response surface model between the end-wall
fence position parameters and the isentropic expansion efficiency is constructed. Additional sampling points are added until the fitting
degree of the response surface meets the requirements. Finally, the predicted values and actual values of the response surface are
shown in Fig. 5(b), which shows a good degree of fitting, and the errors can satisfy subsequent optimization. The optimization results of
the genetic algorithm are shown in Table 2, and the convergence process is shown in Fig. 5(c). The table indicates that candidate point
2 has the best optimization results, and the optimized efficiency reaches 96.162% after verification.
⎧ 2

⎪ H + V 2 ≥ 1352

⎪ H 2 + V 2 ≤ 1452

arctan(H/V) ≤ A + 15 (20)



⎪ arctan(H/V) ≥ A + 5

max η

9
X. Han et al.
Table 2
Optimization results.

Candidate Point 1 Candidate Point 1 (verified) Candidate Point 2 Candidate Point 2 (verified) Candidate Point 3 Candidate Point 3 (verified) ORI
10

A(degree) 57.136 57.136 56.482 56.482 56.073 56.073 –


H (mm) 129.73 129.73 128.61 128.61 129.70 129.70 –
V (mm) 64.495 64.495 66.934 66.934 64.361 64.361 –
η 96.136% 96.131% 96.128% 96.162% 96.123% 96.101% 95.402%

Case Studies in Thermal Engineering 49 (2023) 103306


X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 6. Mach number distribution cloud maps.

In the context of the current vigorous development of new energy, to accommodate new energy, thermal power units need to
improve their ability to deeply adjust peaks to ensure the safe, stable, and economical operation of the power system. Under the
condition of deep peak adjustment, the deviation of each stage of the steam turbine from the design condition will increase, resulting in
increased flow instability and a significant reduction in operating efficiency. To study the effect of the end-wall fence under low-load
conditions, the attack angle generated by the deviation from the design condition was simulated by different inflow angles. To this end,
a comparative study was conducted on flows with inflow angles of 0◦ , 10◦ , and 45◦ , corresponding to the design condition, a slight
deviation from the design condition, and a significant deviation from the design condition, respectively.
The simulations of the previous three cases under the condition of a 45◦ inflow angle were compared with the ORI of the Mach
number cloud map, as shown in Fig. 6. It can be seen from the ORI that a large low-speed area is produced on the pressure surface of the
larger inflow angle, which makes the expansion insufficient and the Mach number before the shock wave and the overall Mach number
decrease. After adding the end-wall fence, the low-speed area in the flow channel is greatly reduced, and the Mach number in case 1 is
almost the same as that in ORI, indicating that the end-wall fence plays a huge role in reducing transverse flow and improving flow
efficiency. For case 2 and case 3, the 45◦ inflow angle is not very different from the design condition, and the end-wall fence greatly
weakens the flow loss caused by the large inflow angle, so it can play a greater role in low-load conditions with a significant deviation
from the design condition.
In previous research, it was found that when the fence position is in the tail region of the flow channel, the improvement in flow
efficiency is greater. Therefore, in the study of multiple inflow angles, the constraint range was narrowed down to expression (21), and
the fence position parameters were slightly modified to obtain more accurate optimization results. The fence position was determined
by three parameters: R, A, and θ, as shown in Fig. 7(a). Fig. 7(b) shows that after changing the constraint conditions, the predicted
values are highly consistent with the actual values, greatly increasing the accuracy of response surface prediction and improving the
subsequent optimization accuracy.


⎪ 130 ≤ R ≤ 145

⎪ 20 ≤ θ ≤ 30



55 ≤ A ≤ 70
(21)

⎪ max η0



⎪ max η10

max η45

After optimization using the multiobjective genetic algorithm, the results are shown in Table 3, and it was confirmed that candidate
point 3 had the best optimization effect. Under the design working condition where the incoming flow angle does not deviate, this
optimized situation is almost indistinguishable from the single-objective optimization results. Moreover, after optimization, the

11
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 7. Optimization scope update and its accuracy.

efficiency was improved by 0.68% for the condition with an incoming flow angle of 10◦ and by 1.67% for the condition with an
incoming flow angle of 45◦ when deviating from the design working condition. This implies that the effectiveness of the end-wall fence
increases as the deviation from the design working condition becomes greater.
The optimized operating condition can be named OPT, and it can be compared with the original operating condition (ORI). The
streamline and shockwave shadowgraph in the flow field are shown in Fig. 8(a) and (b). It was found that the guiding effect of the end-
wall fence on the streamlines is more significant when deviating from the original design condition. The optimized end-wall fence can
greatly reduce the vortices generated in the channel, as clearly seen in the streamline of the 45◦ inflow condition. Moreover, the
shockwaves at the end of the end-wall fence and the stator blades overlap in the shadowgraph, and the end-wall fence breaks the
original shockwave, thereby reducing the shockwave losses. In light of these two aspects, the end-wall fence plays a significant role in
the optimization of the flow structure.

12
X. Han et al.
Table 3
Optimization results of multiple inflow angles.

Candidate Point 1 Candidate Point 1 (verified) Candidate Point 2 Candidate Point 2 (verified) Candidate Point 3 Candidate Point 3 (verified) ORI

θ(degree) 25.608 25.608 26.268 26.268 56.073 56.073 –


13

R(mm) 143.74 143.74 142.47 142.47 129.70 129.70 –


A(degree) 60.193 60.193 58.424 58.424 64.361 64.361 –
η0 96.081% 96.060% 96.024% 95.910% 96.161% 96.098% 95.402%
η10 96.064% 96.003% 96.180% 95.849% 96.094% 96.050% 95.370%
η45 93.563% 93.521% 93.473% 93.437% 93.867% 93.930% 92.264%

Case Studies in Thermal Engineering 49 (2023) 103306


X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

Fig. 8. Comparison of multi-objective genetic algorithm optimization results.

14
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

5. Conclusion
This study focused on the optimization of the end-wall fence position in the White cascade. By parameterizing the end-wall fence
position and using the response surface method, a mapping relationship between the end-wall fence position and the isentropic
expansion efficiency was obtained. A surrogate model was constructed, and single-objective and multiobjective optimizations were
conducted using a genetic algorithm to obtain the optimal end-wall fence position parameters. The main conclusions are as follows:

(1) The end-wall fence not only reduces secondary flow but also effectively reduces shock losses when it is located at the end of the
channel. Therefore, the best optimization effect on the flow structure is achieved when the end-wall fence is located at the end of
the channel.
(2) When the flow deviates significantly from the design condition, the end-wall fence can significantly reduce the low-speed region
on the pressure side of the blade. Therefore, the end-wall fence can play a greater role in low-load conditions with significant
deviations from the design condition.
(3) After optimization for the design condition, the maximum isentropic expansion efficiency was achieved at H = 128.6 mm, V =
66.93 mm, and A = 56.48◦ , reaching 96.160%.
(4) After optimizing for multiple inflow angles under low-load conditions, the best comprehensive optimization result was obtained
at θ = 56.07◦ , R = 129.7 mm, and A = 64.36◦ . The isentropic expansion efficiencies for inflow angles of 0◦ , 10◦ , and 45◦ were
96.098%, 96.050%, and 93.930%, respectively.
The results of this study provide a reference for turbine flow design.

Author statement
Xu Han: Conceptualization, Methodology, Formal analysis, Writing – original draft. Qiuliang Zhu: Conceptualization, Resources,
Supervision, Writing – review & editing. Jiandong Guan: Validation, Investigation. Zhongwen Liu: Investigation. Bochuan Yao:
Software, Validation. Zhonghe Han: Resources.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

No data was used for the research described in the article.

Acknowledgments
The authors are thankful for the support provided by the National Natural Science Foundation of China (52106042), Graduate
Student Innovation Ability Training Funding Project of Hebei Province, China (grant number CXZZSS2023194), and Natural Science
Foundation of Hebei Province, China (grant number E2020502001). and the Fundamental Research Funds for the Central Universities
of China (grant number 022MS084).

References
[1] T. Choudhary, Sanjay, Novel and optimal integration of SOFC-ICGT hybrid cycle: energy analysis and entropy generation minimization, Int. J. Hydrogen Energy
42 (23) (2017) 15597–15612.
[2] M. Yang, Y.L. Zhou, D. Wang, et al., Thermodynamic cycle analysis and optimization to improve efficiency in a 700 degrees C ultrasupercritical double reheat
system, J. Therm. Anal. Calorim. 141 (1) (2019) 83–94.
[3] D.P. Hanak, A.J. Kolios, C. Biliyok, et al., Probabilistic performance assessment of a coal-fired power plant, Appl. Energy 139 (2015) 350–364.
[4] S. Liu, J. Shen, P.H. Wang, Multiparameter joint optimization based on steam turbine thermal system characteristic reconstruction model, Int. Conf. New Energy
and Future Energy Syst. 354 (2019), 012068.
[5] G.J. Zhang, X.Z. Zhang, F.F. Wang, et al., Numerical investigation of novel dehumidification strategies in nuclearplant steam turbine based on the modified
nucleation model, Int. J. Multiphas. Flow 120 (2019), 103083.
[6] J. Gao, Q. Zheng, H. Zhang, Comparative investigation of tip leakage flow and its effect on stage performance in shrouded and unshrouded turbines, Proc. Inst.
Mech. Eng. G J. Aerosp. Eng. 227 (8) (2013) 1265–1276.
[7] Y. Yang, H.P. Peng, C. Wen, A novel dehumidification strategy to reduce liquid fraction and condensation loss in steam turbines, Entropy 23 (9) (2021) 1225.
[8] G.J. Zhang, X.Z. Zhang, F.F. Wang, et al., Design and optimization of novel dehumidification strategies based on modified nucleation model in three-
dimensional cascade, Energy 187 (2019), 115982.
[9] J. Gao, D. Huo, G. Wang, et al., Advances in axial turbine blade profile aerodynamics, Proc. IME C J. Mech. Eng. Sci. 235 (4) (2020) 652–669.
[10] K.N. Kumar, M. Govardhan, Numerical study of effect of streamwise end wall fences on secondary flow losses in two dimensional turbine rotor cascade, Eng.
Appl. Computat. Fluid Mech. 4 (4) (2010) 580–592.
[11] Y.J. Moon, S.R. Koh, Counterrotating streamwise vortex formation in the turbine cascade with end-wall fence, Comput. Fluid 30 (4) (2001) 473–490.
[12] T. Kawai, S. Shinoki, T. Adachi, Visualization study of three-dimensional flows in a turbine cascade end-wall region, JSME Int. J. 33 (2) (1990) 256–264.
[13] T. Poehler, J. Niewoehner, P. Jeschke, et al., Investigation of nonaxisymmetric end-wall contouring and 3D airfoil design in a 1.5 stage axial turbine part I:
design and novel numerical analysis method, in: Proceedings of the ASME Turbo Expo: Turbine Technical Conference and Exposition, 2014.
[14] H.B. Ding, Y. Zhang, Y.Y. Dong, et al., High-pressure supersonic carbon dioxide (CO2) separation benefiting carbon capture, utilisation and storage (CCUS)
technology, Appl. Energy 339 (2023), 120975.
[15] Y.F. Wu, W. Li, D.R. Sheng, et al., Fault diagnosis method of peak-load-regulation steam turbine based on improved PCA-HKNN artificial neural network, Proc.
Inst. Mech. Eng. O J. Risk Reliab. 235 (6) (2021) 1026–1040.

15
X. Han et al. Case Studies in Thermal Engineering 49 (2023) 103306

[16] L. Lin, B.Q. Xu, S.W. Xia, Multi-angle economic analysis of coal-fired units with plasma ignition and oil injection during deep peak shaving in China, Appl. Sci.-
Basel 9 (24) (2020) 5399.
[17] H.J. Wei, Y.W. Lu, Y.C. Yang, et al., Flexible operation mode of coal-fired power unit coupling with heat storage of extracted reheat steam, J. Therm. Sci. 31 (2)
(2022) 436–447.
[18] X. Han, Q.L. Zhu, J.D. Guan, et al., Research on wet steam condensation flow characteristics of steam turbine last stage under zero output condition, Int. J.
Therm. Sci. 179 (2022), 107691.
[19] L. Wang, T.G. Wang, Y. Luo, Improved nondominated sorting genetic algorithm (NSGA)-II in multiobjective optimization studies of wind turbine blades, Appl.
Math. Mech.-Eng.Edition 32 (6) (2011) 739–748.
[20] Z.Y. Li, H.B. Chen, B. Xu, et al., Hybrid wind turbine towers optimization with a parallel updated particle swarm algorithm, Appl. Sci.-Basel 11 (18) (2021)
8683.
[21] S. Rehman, S.S. Ali, S.A. Khan, Wind farm layout design using cuckoo search algorithms, Appl. Artif. Intell. 32 (9) (2018) 956–978.
[22] G.L. Hou, L.J. Gong, Z.L. Yang, et al., Multiobjective economic model predictive control for gas turbine system based on quantum simultaneous whale
optimization algorithm, Energy Convers. Manag. 207 (2020), 112498.
[23] C. Yan, Z.Y. Yin, F.S. Guo, et al., A newly improved collaborative optimization strategy: application to conceptual multidisciplinary design optimization of a civil
aero-engine, in: Proceedings of the ASME Turbo Expo: Turbine Technical Conference and Exposition, 2017.
[24] M.H. Shojaeefard, S.E. Hosseini, J. Zare, CFD simulation and Pareto-based multiobjective shape optimization of the centrifugal pump inducer applying GMDH
neural network, modified NSGA-II, and TOPSIS, Struct. Multidiscip. Optim. 60 (4) (2019) 1509–1525.
[25] K. Peng, J. Fu, D. Fan, et al., Improved genetic algorithm and its application in parameter optimization for certain aeroengine compressor guide vane regulator,
Int. Conf. Electronics, Commun. Control (2011).
[26] J. Cho, K. Kim, J. Kim, et al., Controlling the secondary flows near end-wall boundary layer fences in a 90 degrees turning duct using approximate optimization
method, J. Mech. Sci. Technol. 25 (8) (2011) 2025–2034.
[27] Y.Z. Chen, L. Yang, J.J. Zhong, Numerical study on end-wall fence with varying geometrical parameters in a highly loaded compressor cascade, Aero. Sci.
Technol. 94 (2019), 105390.
[28] R. Uehara, S. Mizuguchi, K. Kusano, Secondary flow loss reduction method by use of 3D-fence in a gas turbine cascade, Proc. ASME/JSME/KSME Joint Fluids
Eng. Conf. 3B (2019).
[29] X. Han, W. Zeng, Z.H. Han, Investigation of the comprehensive performance of turbine stator cascades with heating end-wall fences, Energy 174 (2019)
1188–1199.
[30] H.B. Ding, Y. Zhang, Y. Yang, et al., A modified Euler‒Lagrange‒Euler-Euler approach for modeling homogeneous and heterogeneous condensing droplets and
films in supersonic flows, Int. J. Heat Mass Tran. 200 (2023), 123537.
[31] H.B. Ding, Y. Zhang, C.Q. Sun, et al., Numerical simulation of supersonic condensation flows using Eulerian‒Lagrangian and Eulerian wall film models, Energy
258 (2022), 124833.
[32] F. Mazzelli, F. Giacomelli, A. Milazzo, CFD modeling of condensing steam ejectors: comparison with an experimental test-case, Int. J. Therm. Sci. 127 (2018)
7–18.
[33] G.J. Zhang, S. Dykas, P. Li, et al., Accurate condensing steam flow modeling in the ejector of the solar-driven refrigeration system, Energy 212 (2020), 118690.
[34] A. Kantrowitz, Nucleation in very rapid vapor expansions, J. Chem. Phys. 19 (9) (1951) 1097–1100.
[35] G. Gyarmathy, F. Lesch, Fog droplet observations in laval nozzles and in an experimental turbine, Archive: Proc. Instit. Mech. Eng., Conf. Proc. 184 (37) (1969)
29–36.
[36] K.D. Chandler, A.J. White, J.B. Young, Unsteady wetness effects in LP steam turbines, Proc. ASME TURBO EXPO (2011) 2265–2273.
[37] C. Wen, Y. Yang, H.B. Ding, et al., Wet steam flow and condensation loss in turbine blade cascades, Appl. Therm. Eng. 189 (2021), 116748.
[38] C. Wen, H.B. Ding, Y. Yang, Numerical simulation of nanodroplet generation of water vapor in high-pressure supersonic flows for the potential of clean natural
gas dehydration, Appl. Therm. Eng. 189 (2021), 116748.
[39] Marcus Mccallum, Roland Hunt, The flow of wet steam in a one-dimensional nozzle, Int. J. Numer. Methods Eng. 44 (12) (1999) 1807–1821.
[40] A.J. White, J.B. Young, P.T. Walters, Experimental validation of condensing flow theory for a stationary cascade of steam turbine blades, Phil. Trans. Math.
Phys. Eng. Sci. 354 (1704) (1996) 59–88.

16

You might also like