You are on page 1of 34

Accepted Manuscript

Research Paper

Application of response surface methodology and desirability approach to in-


vestigate and optimize the jet pump in a thermoacoustic Stirling heat engine

P. Yang, H. Chen, Y.W. Liu

PII: S1359-4311(17)33112-5
DOI: http://dx.doi.org/10.1016/j.applthermaleng.2017.08.077
Reference: ATE 10959

To appear in: Applied Thermal Engineering

Received Date: 7 May 2017


Revised Date: 10 August 2017
Accepted Date: 16 August 2017

Please cite this article as: P. Yang, H. Chen, Y.W. Liu, Application of response surface methodology and desirability
approach to investigate and optimize the jet pump in a thermoacoustic Stirling heat engine, Applied Thermal
Engineering (2017), doi: http://dx.doi.org/10.1016/j.applthermaleng.2017.08.077

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Application of response surface methodology and desirability

approach to investigate and optimize the jet pump in a

thermoacoustic Stirling heat engine

Yang P1,2, Chen H1, Liu Y.W1*


1)
Key Laboratory of Thermo-Fluid Science and Engineering of MOE, School of Energy and

Power Engineering, Xi’an Jiaotong University, Xi’an , Shaanxi 710049, P R China


2)
Mechanical Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, IL

61801, USA

Abstract: A jet pump used in the thermoacoustic engine can improve the performance by

suppressing the Gedeon streaming. In this paper, response surface methodology (RSM) and

desirability approach are used to investigate and optimize the jet pump in a thermoacoustic

Stirling heat engine (TASHE). The position, length, diameter and tapered angle of jet pump are

selected as the designing parameters, and pressure amplitude and time-averaged power dissipation

across the jet pump are adopted as responses. The analysis of variance (ANOVA) is conducted to

identify the significant influences of jet pump's parameters on responses. The regression models

are obtained, and prediction deviations for pressure amplitude and power dissipation are within

±2% and ±5% respectively. In addition, the influences of jet pump’s parameters on responses are

plotted by 3D surface. Eventually, the designing parameters of jet pump are optimized to achieve

both maximal pressure amplitude and minimal power dissipation. The optimal results are

validated by comparing with DeltaEC simulation. Results show that the deviations between

DeltaEC and RSM for pressure amplitude and power dissipation are 1.86% and 2.88%, which

indicates that RSM as an assessment methodology in optimizing jet pump is feasible.

Keywords: jet pump, pressure amplitude, power dissipation, response surface methodology,

desirability approach

*Corresponding author. Tel.: (86)13087588436

Email address: ywliu@xjtu.edu.cn

1. Introduction
1
Thermoacoustic effect can be realized as the coversion of heat and sound energy by the

time-averaged interaction between the thermodynamic and acoustic phenomena. It is thriving

since the linear thermoacoustic theory summarized by Rott [1]. As one kinds of thermoacoustic

devices, thermoacoustic heat engines convert heat to acoustic power provided the temperature

gradient along the stack or the regenerator increases above a critical value. Compared with the

traditional mechanical engine, it shows great advantages of no-moving parts, potential high

efficiency and environmental friendliness, especially utilizing waste heat to drive refrigerators

electric generators [2-3]. Therefore, thermoacoustic heat engines involving with complex thermal

and acoustic characteristics earn extensive attention and have been the subject of much research

[4-5].

However, the nonlinear phenomena, such as harmonic content [6-7], vortex flow [8-10],

sound streaming [11-14], the double-threshold effect [15] and so on would limit the further

improvement of the efficiency in thermoacoustic conversion. It's well known that DC streaming

occuring in the loop brings a mountain of heat from a hot heat exchanger to an ambient heat

exchanger and lower the efficiency. A jet pump proposed by Swift [16] used the asymmetry of

hydrodynamic end effects to suppress Gedeon streaming by narrowing the cross-sectional area in a

thermoacoustic Stirling heat engine (TASHE). Experimental results showed that the Carnot

efficiency was up to 40% after using the jet pump. For better understanding the working

mechanism of jet pump in oscillatory flow, numerous work has been carried out. Barton et.al [17]

analyzed power dissipation and time-average pressure through a sudden area change with different

geometries, which showed the way of decreasing power losses across the sudden area. Petculescu

et.al [18] investigated the nonlinear effects and minor losses of the jet pump and discussed the

influence of different cone half-angles. They found that minor loss coefficients was sensitive to

the smallest curvature radius. The inflow minor loss coefficient increased with increasing taper

angle while the outflow coefficient remained relatively constant. They also mentioned that the

impedance of jet pump must be noted if it was applied to a thermoacoustic system. Oosterhuis

et.al developed a two-dimensional CFD model of jet jump in oscillatory flow and investigated the

vortex formation and flow separation of jet pump [19]. They also gave the design guideline in

2
laminar oscillating flow based on numerical parameter study [20]. Tang et.al studied jet pumps

with rectangle and circular tapered angles and compared their performance using time-averaged

and overall resistance [21]. Above mentioned works took the jet pump apart from thermoacoustic

devices and only focused on the performance of the jet pump under oscillating flow condition. If a

jet pump is applied to a thermoacoustic system, the situation would become more complicated. In

addition, the unpredictability of the self-excited oscillation and the complexity of the oscillating

flow in an actual TASHE brought a challenge for designing jet pump. Otherwise, the contraction

of cross section will cause flow separation and energy losses [22]. It was also pointed that the jet

pump could not only restrain the Gedeon streaming, but also affect the acoustic filed in the

regenerator [23]. In our previous study, response surface methodology (RSM) was used to

investigate the effects of jet pump on Gedeon streaming, and further realized the prediction of

Gedeon streaming using the correlation. Results showed that the performance of the TASHE was

not the best when the Gedeon streaming decreases to close to zero [24]. This indicated that energy

losses through a jet pump deteriorated the performance to some extent even when Gedeon

streaming was suppressed under this condition. Therefore, designing a jet pump not only need to

consider the complicated interaction among parameters, but also make a tradeoff between system

performance improvement and energy losses. In order to take both into account, multi-objective

optimization is applied to guide the optimization and designing of a jet pump.

Multi-objective optimization has been employed extensively to predict product performance

and optimize industrial process [25-28]. The desirability function approach, which always

combines with RSM is one of the most widely used methods in industry for the optimization of

multiple response processes. RSM is used to generate design of experiments (DOE) and analyze

parametrical interaction, while desirability approach searches for the optimal combination for

multi-objective by synthetically considering miscellaneous effects [29-31].

In this paper, RSM and desirability approach are used to investigate and optimize the jet

pump in a TASHE. The major parameters of jet pump include its position, inner diameter, length

and tapered angle. Pressure amplitude and power dissipation across the jet pump are adopted as

responses. First the parametric influences on the responses are analyzed visually using analysis of

3
variance and 3-D surface plot. Then the optimal parameters of the jet pump are achieved to

acquire both maximal pressure amplitude and minimal power dissipation. Finally optimal

parameters are validated to confirm the accuracy of predictive results by desirability approach.

2. Mathematical model

2.1 The geometry of the TASHE with a jet pump

Figure 1 presents the schematic diagram of the TASHE with a jet pump. The TASHE is

composed of a loop and a resonator. The main components are arranged successively, which

include a main ambient heat exchanger, a regenerator, a hot heat exchanger, a hot buffer and a

secondary ambient heat exchanger. The detailed dimensions and material of main parts are given

in our previous work [24] and will not shown once more here.

The jet pump is placed on the upside of the main ambient heat exchanger as shown in Figure

1. Figure 2 shows a detailed three-dimensional diagram of the jet pump and its main structural

parameters [24].

In Fig.2, P denotes the position of the jet pump away from the main ambient exchanger (blue

zone in Figure 2). ds is the diameter of the narrow end of the jet pump, which is defined as the

diameter of the jet pump in the following design. The diameter of the wide opening end (db) can

be calculated by

d b = d s + 2 L tan α (1)

where L and α are the length and tapered angle of the jet pump. When the working gas flows

through a jet pump, the abrupt are change (contraction/expansion) from the connecting pipes of

different cross-sectional areas accompanied by jet flow and pressure drop. It's well known that this

pressure drop can be used to suppress the sound streaming and improve the performance. In our

previous study, the parametric analysis of jet pump on sound streaming had been investigated

using RSM [24]. In this paper, we will focus on optimizing the jet pump considering both the

system performance and its energy losses. According to Swift [16], the time-averaged acoustic

power dissipation across a jet pump is:

4
3
. ρ m U1, jp  a 
E= 2  ( K exp, s + K con ,s ) + ( s ) 2 ( K con ,b + K exp,b )  (2)
3π as  ab 

where the subscript “con” and “exp” represent the working gas flows into and out of the opening

of jet pump. “s” and “b” refer to the small and big opening of jet pump. K is the local resistance

loss coefficient at each end of jet pump.

The “Design Environment for Low Amplitude Thermoacoustic Energy Conversion”

(DeltaEC) developed by Swift is used to model and predict the performance of the TASHE [32].

DeltaEC had been validated that it was very effective software and could successfully predict the

performance of thermoacoustic systems and help the design of thermoacoustic system [33, 34].

DelatEC numerically integrates the one-dimensional wave equations including linearized

continuity, momentum and unified energy equation and also combined with gas equation. The

first-order form equations ready for numerical integration are as follows [32]:

dp1 iωρ m
=− U1 (3)
dx A(1 − f v )

dU 1 iω A (γ − 1) fκ ( fκ − f v ) dTm
=− (1 + ) p1 + β U1 (4)
dx ρma 2
1+ εs (1 − σ )(1 + ε s ) dx
where p1 and U1 are pressure amplitude and volume flow rate, ρm and Tm are mean density and

temperature. fκ and fν depend on the structure of tube and the gas properties. a is the sound speed.

β, γ, σ and εs represent thermal expansion coefficient, the ratio of isobaric to isochoric specific

heats, Prandtl number and correction factor for finite solid heat capacity, respectively.

The DeltaEC model has been clarified in detail in our previous study [24], and the operating

conditions and geometrical parameters are listed in table 1 in this article. The minor loss of jet

pump is replaced by a lumped element----MINOR, which accounts conveniently for minor loss

effects (k+ and k-). The minor loss can be used to generate additional pressure drop to suppress

the Gedeon streaming, thus affecting thermoacoustic conversion.

2.2 Validation of DeltaEC model

In order to verify the DeltaEC model, helium gas was used and the condition was the

operating pressure of 1.5 MPa and 2000 W heating power, which remained constant as Yu used
5
[35]. The variation of the pressure amplitude at position P1 (which is at the top of the ambient heat

exchanger near the pressure antinode) along with heating temperature is obtained and shown in

figure 3. In addition, the corresponding experimental results by Yu et al. are also shown in figure

3.

In Figure 3, firstly it can be seen that pressure amplitude increases with heating temperature

both in our model and experimental results. It indicates that the modeling results predict the

thermoacoustic performance qualitatively. Comparing DeltaEC model with Yu’s data, deviation

between them are small when heating temperature is less than 750 K. However, an increasing

deviation is generated as heating temperature continues to rise. When the heating temperature

doesn’t exceed 750 K, the ratio of pressure amplitude to the mean pressure (1.5MPa) is less than

10%. Under this circumstance, the non-linear effects in the thermoacoustic system are negligible,

and the linear thermoacoustic theory in DeltaEC can be used to exactly predict and analyze its

performance. When the heating temperature is higher than 750K, the ratio is higher than 10% and

gradually increases. Under these conditions, the non-linear effect becomes more and more

prominent, and the deviation becomes bigger but no more than 30%. Yu’s experiments include

some non-linear phenomena, such as viscous dissipations and higher order harmonic effect, which

would deteriorate the performance. While DeltaEC model based on linear thermoacoustic theory

loses sight of the non-linear effect. As a result, the DeltaEC simulation prediction is higher than

the experimental data. However, the DeltaEC results still can be used to predict the performance

qualitatively. Totally, the modeling results predict the thermoacoustic performance both

qualitatively and quantitatively.

3. Response surface methodology (RSM) and desirability approach

This paper is devoted to analyze the effect of the jet pump’s parameters on the performance

of the TASHE and then optimize the jet pump. The RSM is employed to design the simulation

experiments and achieve the prediction model for the TASHE with the jet pump. One important

step is to establish a suitable approximation of the functional relationship between the designing

parameters and the responses through RSM. Considering the interactive effect between parameters,

a quadratic regression model is given as follows [36]:


6
k k −1 k k

y = λ0 + ∑ λi xi + ∑ ∑ λij xi x j + ∑ λii xi + ε
2
(5)
i =1 i =1 j =i +1 i =1

where y is the response of the system (Gedeon streaming), xi and xj coded independent variables; k

is the number of variables; λ0, λi , λii and βij (i=0,1,2,…, k; j=i+1,1,2,…,k) are the regression

coefficients for the intercept, linear, quadratic and interaction terms respectively; and ε is the

statistical error.

The Box-Behnken design (BBD) is performed to arrange simulation experiments because it

accomplishes the task with the non-sequential analysis and fewer design points. Four factors and

three levels (-1, 0, 1) are used for the construction of a second-order response surface model. The

variables used in the study are the position, length, diameter, and tapered angle of the jet pump.

The range of variables is determined by the actual design structure [16,18]. In addition, some

additional points from central composite design are added to ensure the accuracy of the

second-order model. The variable values generated from the RSM are used for numerical solutions

to obtain pressure amplitude and time-averaged power dissipation across the jet pump. Pressure

amplitude represents the intensity of acoustic field. The time-averaged power dissipation across

the jet pump evaluates the energy loss. It is expected to be as small as possible. The aim is to

optimize the parameters of jet pump, which can achieve maximal pressure amplitude and minimal

power dissipation. The ranges and levels of variables was given in [24] and will not shown here.

The desirability method is employed to realize the multi-objective optimization due to its

simplicity and flexibility for each response. Each response is transformed to a dimensionless

desirability value ranging from 0 to 1. 0 indicates the response is outside of the limits and

unacceptable while 1 suggests the response is the most desirable. For several responses and factors,

all goals get combined into one desirability function. The simultaneous objective function is a

geometric mean of all transformed responses [36]:

n ri

D = (d1r1 × d 2 r2 × ... × d n rn ) ∑ i = (∏ d i ) ∑ i
1/ r 1/ r
(6)
i =1

where n is the number of responses in the measure. If any of the responses or factors fall outside

their desirability range, the overall function becomes zero. For simultaneous optimization each
7
response must have a low and high value assigned to each goal. On the worksheet, the "Goal"

field for responses must be one of five choices: "none", "maximum", "minimum", "target", or "in

range". Factors will always be included in the optimization, at their design range by default, or as

a maximum, minimum of target goal. The meanings of the goal parameters are defined as

following.

For goal of maximum, the desirability will be defined by the following equation:

 di = 0, Yi ≤ Low i
 wt i
  Yi − Low i 
 di =   , Low i < Yi < High i (7)
  High i − Low i 
 d = 1, Y ≥ High
 i i i

For goal of minimum, the desirability will be defined by the following equation:

 di = 1, Yi ≤ Low i
 wt i
  High i − Yi 
 di =   , Low i < Yi < High i (8)
  High i − Low i 
 d = 0, Y ≥ High
 i i i

For goal within range, the desirability will be defined by the following equation:

 di = 0, Yi < Low i

 di = 1, Low i ≤ Yi ≤ High i (9)
 d = 0, Y > High
 i i i

The numerical optimization finds a point that maximizes the desirability function. In this paper,

the goals of maximum and minimum are applied to determine individual desirability values for

pressure amplitude and time-averaged power dissipation through the jet pump, respectively.

4. Results and discussion

According to the design of experiment, each run with different factor groups is considered as

an input of DeltaEC model to obtain the corresponding response. The design of experimental

matrix including factors and responses is presented in Table 2. A, B, C and D represent the

position, length, diameter and tapered angle of the jet pump respectively. Except for pressure

amplitude and power dissipation, the temperature at the hot heat exchanger and minor loss

8
coefficients are also listed in table 2. In order to clarify the variation of pressure amplitude, the

Gedeon streaming from our previous study [24] is also shown in table 2. It can be seen that the

variation of pressure amplitude is due to the combined effect of the temperature at the hot heat

exchanger and Gedeon streaming.

4.1 Analysis of variance (ANOVA)

In the following, the statistical analysis of the responses will be done by ANOVA. In ANOVA,

the sum of squares is used to estimate the square of deviation from the grand mean. Mean squares

are estimated by dividing the sum of squares by degrees of freedom. The F-ratio is defined as the

ratio of mean square for the term to the mean square for the residual. It is used to check the

accuracy of the model in which the calculated value of F should be greater than the F-table value.

If the p-values in the ANOVA table are less than 0.05, then the selected factors have significant

effects on the responses.

Table 3 shows ANOVA results for pressure amplitude. “Adeq Precision” measures the signal

to noise ratio. A ratio greater than 4 is desirable. For pressure amplitude, the ratio of 37.470

indicates an adequate signal. Therefore, this model can be used to navigate the design space. The

“R-Squared” of 0.9683 indicates the model has a large value of goodness-of-fit. Values of "Prob >

F" less than 0.05 imply model terms are significant. In this case, all terms except for interaction

term BD have significant effects on pressure amplitude.

Table 4 shows ANOVA results for time averaged power dissipation. The “Adeq Precision” of

34.174 indicates this model can be used to navigate the design space. The model has a large value

of goodness-of-fit (R-Squared= 0.9669). In term of p-values less than 0.05, all terms except for the

quadratic term A2 have significant influence on time averaged power dissipation.

4.2 Regression model of responses

After the significance of factors is identified by ANOVA, the quadratic regression models for

pressure amplitude and time averaged power dissipation within the limits of the factors are

obtained as follows:

9
-3 -3 -3 -5
p1 = −0.4411 − 6.81 × 10 × A − 4.38 × 10 × B + 0.1180 × C − 2.08 × 10 × D − 1.92 × 10 × A × B
-5 -5 -4 -5 -3
+5.38 × 10 × A × C − 8.01 × 10 × A × D + 2.34 × 10 × B × C +1.70 × 10 × B × D +1.19 × 10 × C × D (10)
-5 2 -6 2 -3 2 -4 2
−1.81 × 10 × A +7.92 × 10 × B − 4.56 × 10 × C +1.05 × 10 × D

E jp = − 476.73 − 3.08 × A − 4.35 × B + 95.98 × C − 21.04 × D − 0.02 × A × B + 0.34 × A × C


− 0.08 × A × D + 0.13 × B × C + 0.06 × B × D + 0.80 × C × D − 0.01 × A + 0.02 × B
2 2
(11)
2 2
− 3.59 × C + 0.27 × D

The model obtained above can be used to predict pressure amplitude and time-averaged

power dissipation within the given range the factors. The deviations between pressure amplitude

and time-averaged power dissipation obtained from DeltaEC and predicted by the regression

model are illustrated in Figure 4. It can be seen that most of design points are distributed close to

the diagonal line and deviations between the actual (DeltaEC) and the predicted (regression model)

values for pressure amplitude, time-averaged power dissipation and relative Carnot efficiency are

almost within ±2% and ±5%, respectively. It implies the regression model can be used to predict

performance accurately.

4.3 Influence of jet pump’s parameters

Fig.5 presents the interactive effect of the position, diameter, length and tapered angle on

pressure amplitude. Fig.5(a) shows the combined effects of the position, diameter on pressure

amplitude while keeping two other factors constant by 3-D surface plot. It can be seen that the

interactive influence is complex. For a small diameter, pressure amplitude decreases with the

position increasing. As the diameter increases, it remains relatively constant. The maximal

pressure amplitude can be achieved by decreasing position and with diameter of 17-18 mm. The

combined influence of length and tapered angle on pressure amplitude is presented in Figure 5(b).

It shows that the small length and tapered angle can increase pressure amplitude while fixing the

position at 35 mm and diameter at 16.5 mm.

Fig.6 presents the interactive effect of the position, diameter, length and tapered angle on the

time-averaged power dissipation across the jet pump. For a small diameter, power dissipation

remains relatively constant with the variation of the position. For a large diameter, it increases

with the position. When position is small, the power dissipation first increases and then decreases

with the diameter. While it increases with the diameter for the large position. Fixing length of 40
10
mm and tapered angle 10°, the minimal power dissipation can be achieved by position of 20mm

and diameter of 18 mm. Fig 6(b) shows that the power dissipation first reduces then increases

slightly with an increase of length and tapered angle. The large length and tapered angle are

beneficial to decrease power dissipation.

4.4 Multi-objective optimization

The goal of the parametric analysis for the jet pump is to find the optimized designing

parameters of the jet pump where pressure amplitude can be higher and power dissipation can be

lower. Desirability approach is used to carried out the multi-objective optimization. The position,

length, diameter and tapered angle of jet pump vary in the design range. Pressure amplitude and

power dissipation are set to be maximal and minimal. Equal importance and weight are assigned

for all variables and objectives. Table 5 shows the goal, lower and upper limits and weight as well

as the importance for each factor and response.

Table 6 shows the optimal parameters of jet pump where the position, length, inner diameter

and tapered angle are 20 mm, 51.59 mm, 18.00 mm and 13.45°, respectively. The corresponding

pressure amplitude and power dissipation is 0.2632 MPa and 56.04 W. The combined desirability

is 0.9298 which is closed to 1. That indicates the optimal results are more desirable.

To validate optimal results, the same parameters are also input into DeltaEC model, and the

results are also shown in Table 7. Compared to the results between RSM and DeltaEC with

optimized jet pump, the deviation of pressure amplitude and power dissipation are 1.86% and

2.88%, respectively. It shows that the optimal results predicted by RSM and desirability approach

agree well with DeltaEC results.

5. Conclusion

In this study, an effective procedure of response surface methodology (RSM) and desirability

approach has been developed to carry out parametric analysis and optimization of the jet pump.

The main conclusions can be drawn as follows.

(1) The significant influences of jet pump’s parameters on pressure amplitude and power

dissipation are tested statistically by ANOVA. High “Adeq Precision” value and “R-Squared”

value indicates the model has a better goodness-of-fit and can navigate the design.
11
(2) The RSM regression models for pressure amplitude and power dissipation are obtained. It is

found that the deviations between prediction results by RSM and actual values by DeltaEC are

within ±2% and ±5% for pressure amplitude and power dissipation respectively.

(3) The interactive influences of the jet pump’s parameters on pressure amplitude and time

averaged power dissipation are illustrated and analyzed in 3D surface plot. Increasing the diameter

and decreasing position of jet pump can improve pressure amplitude and reduce power dissipation.

However, the small length and tapered angle increase both pressure amplitude and power

dissipation. There is a trade-off between pressure amplitude and power dissipation when changing

length and tapered angle of the jet pump.

(4) The optimized parameters of the jet pump are obtained to achieve both maximal pressure

amplitude and minimal power dissipation using desirability approach. Compared to the predicted

values by RSM and calculated results by DeltaEC with the optimized jet pump, the deviations of

pressure amplitude and power dissipation are 1.86% and 2.88%. It shows that RSM and

desirability approach are reliable to optimize the jet pump.

Acknowledgement

This work is supported by the Chinese National Natural Science Foundation (No.51576150).

Nomenclature

A cross-sectional area (m2) p1 pressure amplitude (Pa)

a sound speed (m/s) T temperature (K)

d diameter of jet pump(m) U1 volumetric velocity (m3/s)

E time-averaged power dissipation (W) Q input heat (W)

f spatially averaged diffusion function W acoustic power (W)

K local resistance coefficient x variables

L length of jet pump (m) y response

Greek symbols

α tapered angle of jet pump (°) ρ density (kg/m3)

β thermal expansion coefficient (K-1) ε statistical error

γ the ratio of isobaric to isochoric specific heats σ Prandtl number

12
η relative Carnot efficiency ω angular frequency

εs correction factor for finite solid heat capacity λ regression coefficient

Subscripts

b narrow end of jet pump m mean

c cold end of the regenerator h hot end of the regenerator

k number of factors s wide end of jet pump

con flow into the opening of jet pump

exp flow out of the opening of jet pump

Other

|| amplitude

Reference

[1] Rott N. Thermoacoustics[J]. Archive of Applied Mechanics, 1980, 20: 135-175.

[2] Jin T, Huang J, Feng Y, et al. Thermoacoustic prime movers and refrigerators: Thermally

powered engines without moving components[J]. Energy, 2015, 93: 828-853.

[3] Sun D M, Wang K, Zhang X J, et al. A traveling-wave thermoacoustic electric generator with

a variable electric RC load[J]. Applied Energy, 2013, 106: 377-382.

[4] Kamsanam W, Mao X, Jaworski A J. Thermal performance of finned-tube thermoacoustic

heat exchangers in oscillatory flow conditions[J]. International Journal of Thermal Sciences, 2016,

101: 169-180.

[5] Shi L, Yu Z, Jaworski A J. Application of laser-based instrumentation for measurement of

time-resolved temperature and velocity fields in the thermoacoustic system[J]. International

Journal of Thermal Sciences, 2010, 49(9): 1688-1701.

[6] Swift G W. Analysis and performance of a large thermoacoustic engine[J]. The Journal of the

Acoustical Society of America, 1992, 92(3): 1551-1563.

[7] Luo E, Ling H, Dai W et al. A high pressure-ratio, energy-focused thermoacoustic heat

engine with a tapered resonator[J]. Chinese Science Bulletin, 2005, 50(3): 284-286.

[8] Berson A, Blanc-Benon P. Nonperiodicity of the flow within the gap of a thermoacoustic

couple at high amplitudes[J]. The Journal of the Acoustical Society of America, 2007, 122(4):

13
EL122-EL127.

[9] Shi L, Yu Z, Jaworski A J. Vortex shedding flow patterns and their transitions in oscillatory

flows past parallel-plate thermoacoustic stacks[J]. Experimental Thermal and Fluid Science, 2010,

34(7): 954-965.

[10] Yang, P, Chen H, and Liu Y W. Numerical investigation on nonlinear effect and vortex

formation of oscillatory flow throughout a short tube in a thermoacoustic Stirling engine[J].

Journal of Applied Physics, 2017, 121(21): 214902.

[11] Bailliet H, Gusev V, Raspet R, et al. Acoustic streaming in closed thermoacoustic devices[J].

The Journal of the Acoustical Society of America, 2001, 110(4): 1808-1821.

[12] Biwa T, Tashiro Y, Ishigaki M, et al. Measurements of acoustic streaming in a looped-tube

thermoacoustic engine with a jet pump[J]. Journal of applied physics, 2007, 101(6): 064914.

[13] Penelet G, Guedra M, Gusev V, et al. Simplified account of Rayleigh streaming for the

description of nonlinear processes leading to steady state sound in thermoacoustic engines[J].

International Journal of Heat and Mass Transfer, 2012, 55(21): 6042-6053.

[14] Tang K, Feng Y, Jin S H, et al. Performance comparison of jet pumps with rectangular and

circular tapered channels for a loop-structured traveling-wave thermoacoustic engine[J]. Applied

Energy, 2015, 148: 305-313.

[15] [J]. Applied Thermal Engineering, 2017, 115: 1089-1100.

[16] Penelet G, Gaviot E, Gusev V, et al. Experimental investigation of transient nonlinear

phenomena in an annular thermoacoustic prime-mover: observation of a double-threshold

effect[J]. Cryogenics, 2002, 42(9): 527-532.

[17] Backhauss S, Swift G W. A thermoacoustic Stirling heat engine[J]. Nature, 1999, 399(27):

335-338.

[18] Smith B L, Swift G W. Power dissipation and time-averaged pressure in oscillating flow

through a sudden area change[J]. The Journal of the Acoustical Society of America, 2003, 113(5):

2455-2463.

[19] Petculescu A, Wilen L A. Oscillatory flow in jet pumps: Nonlinear effects and minor

losses[J]. The Journal of the Acoustical Society of America, 2003, 113(3): 1282-1292.

14
[20] Oosterhuis J P, Bühler S, van der Meer T H, et al. A numerical investigation on the vortex

formation and flow separation of the oscillatory flow in jet pumps[J]. The Journal of the

Acoustical Society of America, 2015, 137(4): 1722-1731.

[21] Oosterhuis J P, Bühler S, Wilcox D, et al. Jet pumps for thermoacoustic applications: Design

guidelines based on a numerical parameter study [J]. The Journal of the Acoustical Society of

America, 2015, 138(4): 1991-2002.

[22] Tang K, Feng Y, Jin S, et al. Performance comparison of jet pumps with rectangular and

circular tapered channels for a loop-structured traveling-wave thermoacoustic engine[J]. Applied

Energy, 2015, 148: 305-313.

[23] Oosterhuis J P, Verbeek A A, Bühler S, et al. Flow Separation and Turbulence in Jet Pumps

for Thermoacoustic Applications[J]. Flow, turbulence and combustion, 2017, 98(1): 311-326.

[24] Liu Y W, and Yang P. Influence of inner diameter and position of phase adjuster on the

performance of the thermo-acoustic Stirling engine[J]. Applied Thermal Engineering, 2014, 73(1):

1141-1150.

[25] Yang P, Liu Y W and Zhong G Y. Prediction and parametric analysis of acoustic streaming

in a thermoacoustic Stirling heat engine with a jet pump using response surface methodology[J].

Applied Thermal Engineering, 2016, 103: 1004-1013.

[26] Goel T, Vaidyanathan R, Haftka R T, et al. Response surface approximation of Pareto

optimal front in multi-objective optimization[J]. Computer methods in applied mechanics and

engineering, 2007, 196(4): 879-893.

[27] Rout S K, Choudhury B K, Sahoo R K, et al. Multi-objective parametric optimization of

Inertance type pulse tube refrigerator using response surface methodology and non-dominated

sorting genetic algorithm[J]. Cryogenics, 2014, 62: 71-83.

[28] Han H, Li B, and Shao W. Multi-objective optimization of outward convex corrugated tubes

using response surface methodology[J]. Applied Thermal Engineering, 2014, 70(1): 250-262.

[29] Tartibu L K, Sun B, and Kaunda M A E. Multi-objective optimization of the stack of a

thermoacoustic engine using GAMS[J]. Applied Soft Computing, 2015, 28: 30-43.

[30] Alaswad A, Benyounis K Y, and Olabi A G. Employment of finite element analysis and

15
Response Surface Methodology to investigate the geometrical factors in T-type bi-layered tube

hydroforming[J]. Advances in Engineering Software, 2011, 42(11): 917-926.

[31] Najafi G, Ghobadian B, Yusaf T, et al. Optimization of performance and exhaust emission

parameters of a SI (spark ignition) engine with gasoline–ethanol blended fuels using response

surface methodology[J]. Energy, 2015, 90: 1815-1829.

[32] Hirkude J B, Padalkar A S. Performance optimization of CI engine fuelled with waste fried

oil methyl ester-diesel blend using response surface methodology[J]. Fuel, 2014, 119: 266-273.

[33] Ward B, Clark J, Swift G W. Design Environment for Low-amplitude Thermoacoustic

Energy Conversion (DeltaEC Version 6.2) Users Guide[J]. Los Alamos national laboratory, 2008.

[34] Zhang X Q, Chang J Z. Onset and steady-operation features of low temperature differential

multi-stage travelling wave thermoacoustic engines for low grade energy utilization[J]. Energy

Conversion and Management,2015, 105: 810-816.

[35] Chen M, Ju Y L. Effect of different working gases on the performance of a small

thermoacoustic Stirling engine[J]. International Journal of Refrigeration, 2015, 51: 41-51.

[36] Yu G Y, Luo E C, Dai W, et al. Study of nonlinear processes of a large experimental

thermoacoustic-Stirling heat engine by using computational fluid dynamics[J]. Journal of Applied

Physics, 2007, 102(7): 074901.

[37] Myers R H, Montgomery D C and Anderson-Cook C M. Response surface methodology:

process and product optimization using designed experiments. John Wiley & Sons, 2016.

16
17

Table lists:

Table 1 Operational and geometrical parameters in the TASHE with a jet pump

Table 2 Design of experimental matrix

Table 3 ANOVA for pressure amplitude

Table 4 ANOVA for time averaged power dissipation

Table 5 The multi-objective optimization

Table 6 Optimal results based on desirability approach

Table 7 Confirmation of optimal parameters

18
Table 1 Operational and geometrical parameters in the TASHE with a jet pump

Gas Helium

Initial mean pressure 3.0 MPa


Operational condition
Heat input 3000 W

Acoustic load 5e07 Pa.s/m3

Component Diameter (m) Length (m) Material Porosity

Feedback tube 0.080 2.980 Stainless

Main ambient heat exchanger 0.080 0.045 Copper plate Porous zone 0.5

Regenerator 0.080 0.080 Stainless screen Porous zone 0.697

Hot heat exchanger 0.080 0.058 Copper plate Porous zone 0.806

Thermal buffer tube 0.080 0.240 Stainless

Secondary ambient heat


0.080 0.022 Copper plate Porous zone 0.5
exchanger

Resonator 0.080-0.200 5.000 Stainless

19
Table 2 Design of experimental matrix

Factors Responses

Pressure Power Gedeon


Run A B C D Temperature
amplitude dissipation K+ K- streaming
(mm) (mm) (mm) (°) (K)
(MPa) (W) (mol/s)

1 20 20 16.5 10 0.2672 79.2 1.060 0.701 718 -0.0008


2 50 20 16.5 10 0.259 109.2 1.060 0.701 820 0.0637
3 20 60 16.5 10 0.2611 59 0.962 0.514 711 0.0461
4 50 60 16.5 10 0.2279 66 0.962 0.514 743 0.1114
5 35 40 15 5 0.2532 88.6 1.058 0.684 844 0.0793
6 35 40 18 5 0.2499 77.8 1.062 0.720 666.8 -0.035
7 35 40 15 15 0.226 52 0.968 0.510 822 0.1336
8 35 40 18 15 0.2595 66.8 0.955 0.515 686 0.0317
9 20 40 16.5 5 0.263 76.6 1.061 0.703 708 -0.0083
10 50 40 16.5 5 0.2574 109.6 1.061 0.703 794 0.057
11 20 40 16.5 15 0.2624 59 0.962 0.513 718 0.05
12 50 40 16.5 15 0.2316 66.6 0.962 0.513 766 0.1254
13 35 20 15 10 0.2523 86 1.057 0.682 871 0.0854
14 35 60 15 10 0.2256 52.8 0.968 0.511 798 0.1342
15 35 20 18 10 0.2552 82 1.061 0.718 675 -0.0264
16 35 60 18 10 0.2576 66 0.955 0.516 680 0.0265
17 20 40 15 10 0.2524 59 0.988 0.548 782 0.0898
18 50 40 15 10 0.2022 52 0.988 0.548 896 0.1474
19 20 40 18 10 0.2561 57.3 0.981 0.563 655 -0.0165
20 50 40 18 10 0.2521 80.3 0.981 0.563 712 0.0492
21 35 60 16.5 5 0.2579 79.7 1.011 0.606 738 0.0501
22 35 20 16.5 15 0.2608 79.6 1.009 0.602 764 0.0621
23 35 60 16.5 15 0.2474 62.8 0.951 0.494 731 0.0931
24 27.5 30 15.75 7.5 0.266 83.6 1.044 0.663 782 0.0473
25 42.5 30 15.75 7.5 0.2514 91.5 1.044 0.663 823 0.0829
26 27.5 50 15.75 7.5 0.2548 67.4 0.991 0.562 759 0.0761
27 42.5 50 15.75 7.5 0.2341 70.2 0.991 0.562 770 0.1174
28 27.5 30 17.25 7.5 0.2613 78.6 1.045 0.679 694 -0.0074
29 42.5 30 17.25 7.5 0.2631 96.8 1.045 0.679 735 0.0249
30 27.5 50 17.25 7.5 0.2612 68.6 0.988 0.570 696 0.0207
31 42.5 50 17.25 7.5 0.2553 79.6 0.988 0.570 723 0.0567
32 27.5 30 15.75 12.5 0.2555 66.8 0.991 0.560 771 0.0803
33 42.5 30 15.75 12.5 0.2336 68.5 0.991 0.560 786 0.1211
34 27.5 50 15.75 12.5 0.2484 59.5 0.964 0.509 752 0.0951
35 42.5 50 15.75 12.5 0.2279 59.5 0.964 0.509 778 0.1323

20
36 27.5 30 17.25 12.5 0.2635 69.4 0.987 0.569 703 0.0267
37 42.5 30 17.25 12.5 0.2568 79.6 0.987 0.569 735 0.0625
38 27.5 50 17.25 12.5 0.2602 62.8 0.957 0.511 698 0.0494
39 42.5 50 17.25 12.5 0.2505 70.6 0.957 0.511 721 0.0783
40 20 40 16.5 10 0.2643 63.8 0.985 0.556 717 0.0366
41 50 40 16.5 10 0.2364 75.9 0.985 0.556 760 0.1129
42 35 20 16.5 10 0.2678 96 1.060 0.701 766 0.0303
43 35 60 16.5 10 0.2495 65.7 0.962 0.514 732 0.0845
44 35 40 15 10 0.2295 58.6 0.988 0.548 809 0.1276
45 35 40 18 10 0.2596 71.4 0.981 0.563 683 0.0156
46 35 40 16.5 5 0.2647 94.7 1.061 0.703 749 0.0226
47 35 40 16.5 15 0.2502 65.3 0.962 0.513 741 0.0881
48 35 40 16.5 10 0.2548 72 0.985 0.556 745 0.073

21
Table 3 ANOVA for pressure amplitude

Sum of Mean p-value


Source df F - Value Significant
Squares Square Prob > F
Model 0.008794 14 0.000628 71.9767 < 0.0001 *
A 0.002427 1 0.002427 278.0795 < 0.0001 *
B 0.001003 1 0.001003 114.9551 < 0.0001 *
C 0.002198 1 0.002198 251.8523 < 0.0001 *
D 0.000609 1 0.000609 69.8416 < 0.0001 *
AB 0.000166 1 0.000166 19.0753 0.0001 *
AC 0.000732 1 0.000732 83.8860 < 0.0001 *
AD 0.000181 1 0.000181 20.6951 < 0.0001 *
BC 0.000246 1 0.000246 28.1550 < 0.0001 *
BD 1.06E-05 1 1.06E-05 1.2182 0.2777
CD 0.000397 1 0.000397 45.4856 < 0.0001 *
A2 0.000114 1 0.000114 13.1110 0.0010 *
B2 6.3E-05 1 6.3E-05 7.2226 0.0112 *
C2 0.000724 1 0.000724 83.0023 < 0.0001 *
D2 4.37E-05 1 4.37E-05 5.0086 0.0321 *
Residual 0.000288 33 8.73E-06
Cor Total 0.009082 47
Adeq Precision =37.470
R-Squared=0.9683 Adj R-Squared=0.9548

22
Table 4 ANOVA for time averaged power dissipation

Sum of Mean p-value


Source df F - Value Significant
Squares Square Prob > F
Model 8630.757 14 616.4826 68.9084 < 0.0001 *
A 1020.014 1 1020.014 114.0138 < 0.0001 *
B 3119.059 1 3119.059 348.6382 < 0.0001 *
C 288.8006 1 288.8006 32.2812 < 0.0001 *
D 3029.397 1 3029.397 338.6161 < 0.0001 *
AB 146.882 1 146.882 16.4180 0.0003 *
AC 299.538 1 299.538 33.4814 < 0.0001 *
AD 185.4405 1 185.4405 20.7279 < 0.0001 *
BC 79.6005 1 79.6005 8.8975 0.0053 *
BD 150.5792 1 150.5792 16.8313 0.0003 *
CD 178.802 1 178.802 19.9859 < 0.0001 *
A2 22.23062 1 22.23062 2.4849 0.1245
B2 333.1791 1 333.1791 37.2417 < 0.0001 *
C2 449.182 1 449.182 50.2081 < 0.0001 *
D2 289.8693 1 289.8693 32.4006 < 0.0001 *
Residual 295.2314 33 8.946406
Cor Total 8925.988 47
Adeq Precision =34.174
R-Squared=0.9669 Adj R-Squared=0.9529

23
Table 5 The multi-objective optimization

Lower Upper Lower Upper


Name Goal Importance
Limit Limit Weight Weight

A: Position is in range 20 50 1 1 3
B: Length is in range 20 60 1 1 3
C: Diameter is in range 15 18 1 1 3
D: Tapered angle is in range 5 15 1 1 3
Pressure amplitude maximize 0.2022 0.2678 1 1 3
Power dissipation minimize 52 109.6 1 1 3

24
Table 6 Optimal results based on desirability approach

Optimal value Desirability


Position 20 mm 1
Length 51.59 mm 1
Diameter 18.00 mm 1
Tapered angle 13.45° 1
Pressure amplitude 0.2632 MPa 0.8040
Power dissipation 56.04 W 0.8040
Combined - 0.9298

25
Table 7 Confirmation of optimal parameters

Position Length Diameter Tapered Pressure Power


No.
(mm) (mm) (mm) angle (°) amplitude dissipation
1 20 51.59 18.00 13.45 Predicted 0.2632 56.04
DeltaEC 0.2584 54.47
Error 1.86% 2.88%

26
Figure lists:

Fig.1 The schematic illustration of the TASHE with a jet pump

Fig.2 The schematic illustration of the jet pump

Fig.3 The comparison of thermoacoustic performance by DeltaEC and Yu et al. [35]

Fig.4 Actual (DeltaEC) versus predicted values by RSM

Fig.5 The response surface 3D plot of pressure amplitude

Fig.6 The response surface 3D plot of time-averaged power dissipation

27
Fig.1 The schematic illustration of the TASHE with a jet pump

28
(a) (b)

Fig.2 The schematic illustration of the jet pump

29
0.25
Our results

Pressure amplitude at position P1 (MPa)


Experiment data by Yu et al.
0.20

0.15

0.10

0.05

0.00
550 600 650 700 750 800 850 900
Heating temperature (K)
Fig.3 The comparison of thermoacoustic performance by DeltaEC and Yu et al. [35]

30
(a) Pressure amplitude (b) Power dissipation
Fig.4 Actual (DeltaEC) versus predicted values by RSM

31
(a) Interactive effects of position and diameter

(b) Interactive effects of length and tapered angle

Fig.5 The response surface 3D plot of pressure amplitude

32
(a) Interactive effects of position and diameter

(b) Interactive effects of length and tapered angle

Fig.6 The response surface 3D plot of time-averaged power dissipation

33

You might also like