• No results found

Acoustic Properties of an In-Duct Orifice Subjected to Bias Flow and High Level Acoustic Excitation

N/A
N/A
Protected

Academic year: 2021

Share "Acoustic Properties of an In-Duct Orifice Subjected to Bias Flow and High Level Acoustic Excitation"

Copied!
8
0
0

Loading.... (view fulltext now)

Full text

(1)

http://www.diva-portal.org

Postprint

This is the accepted version of a paper presented at 10th International Conference on Flow-Induced Vibration (& Flow-Induced Noise); Dublin, Ireland, 2nd - 6th July 2012.

Citation for the original published paper:

Bodén, H., Zhou, L. (2012)

Acoustic Properties of an In-Duct Orifice Subjected to Bias Flow and High Level Acoustic Excitation.

In: (ed.), Proceedings of the 10th International conference on Flow-Induced Vibration (& Flow- Induced Noise): FIV2012, Dublin, Ireland, 3-6 July 2012 FLOW-INDUCED VIBRATION (pp.

187-193).

N.B. When citing this work, cite the original published paper.

Permanent link to this version:

http://urn.kb.se/resolve?urn=urn:nbn:se:kth:diva-128455

(2)

ACOUSTIC PROPERTIES OF AN IN-DUCT ORIFICE SUBJECTED TO BIAS FLOW AND HIGH LEVEL ACOUSTIC EXCITATION

Hans Bod´en Linn´eFlow Centre Marcus Wallenberg Laboratory Aeronautical and Vehicle Engineering KTH -The Royal Institute of Technology

MWL,AVE,KTH,Stockholm,SE-10044 Email: hansbod@kth.se

Lin Zhou

CCGEx-Competence Centre for IC-Engine Gas Exchange Marcus Wallenberg Laboratory

Aeronautical and Vehicle Engineering KTH -The Royal Institute of Technology

MWL,AVE,KTH,Stockholm,SE-10044 Email: linzhou@kth.se

ABSTRACT

This paper experimentally investigates the acoustic properties of an orifice with bias flow under medium and high sound level excitation. The test included no bias flow and two bias speeds for three different frequencies. Ex- perimental results are compared and discussed with the- ory. It is shown that bias flow makes the acoustic prop- erties much more complex compared theory and with the no bias flow case, especially when velocity ratio between acoustic particle velocity and mean flow velocity is near unity.

NOMENCLATURE c Speed of sound Cc Discharge coeffecient d Orifice diameter D Diameter of pipe J Bessel function

l Variable effective length lw Orifice thickness

l0 End correction on one side of orifice L Jet length

Mb Bias flow Mach number Mg Grazing flow Mach number k Wave number

KR Rayleigh conductivity

U Mean flow velocity in the orifice P Sound pressure amplitude P0 Mean flow pressure drop dP Sound pressure difference

∆P Measured mean flow pressure drop R Orifice radius

Address all correspondence to this author.

U Mean flow velocity in the orifice V Acoustic particle velocity in the hole ZR Acoustic resistance

ZI Acoustic reactance Z Acoustic impedance β =d/D

µ Adiabatic dynamic viscosity ν Kinematic viscosity

ω Radian frequency ρ0 Air density σ Porosity

τ Time for ejection of air flow orifice Superscripts

ˆ Denotes a peak value Subscripts

u On source side of orifice d On downside of orifice + Mean flow direction

− Opposite direction of mean flow 1, 2 Microphones on source side

3, 4 Microphones on the downstream side

INTRODUCTION

Orifice plates and perforates appear in many technical applications where they are exposed to a combination of high acoustic excitation levels and either grazing or bias flow or a combination. Examples are automotive muf- flers and aircraft engine liners. Taken one by one the ef- fect of high acoustic excitation levels, bias flow and graz- ing flow are reasonably well understood. The nonlinear effect of high level acoustic excitation has for instance

(3)

been studied in [1–11]. It is well known from this liter- ature that perforates can become non-linear at fairly low acoustic excitation levels. The non-linear losses are as- sociated with vortex shedding at the outlet side of the orifice or perforate openings [9, 10]. The effect of bias flow has for instance been studied in [12–17]. Losses are significantly increased in the presence of bias flow, since it sweeps away the shed vortices and transforms the ki- netic energy into heat,without further interaction with the acoustic field. Grazing flow has also received a lot of attention see for instance [18–24]. The combination of bias flow and high level acoustic excitation has been dis- cussed and studied in [25] and some experimental inves- tigations have been made in [26]. Luong [25] derived a simple Rayleigh conductivity model for cases when bias flow dominates and no flow reversal occurs.

The purpose of the present paper is to make a detailed experimental study of the transition between the case when high level nonlinear acoustic excitation is the factor determining the acoustic properties to the case when bias flow is most important. As discussed in [25] it can from a theoretical perspective be expected that this is related to if high level acoustic excitation causes flow reversal in the orifice or if the bias flow maintains the flow direc- tion. Acoustic properties, such as impedance, Rayleigh conductivity and absorption coefficient are discussed.

Semi-empirical impedance model

Starting for instance from [4] a number of semi- empirical models have been developed to include the ef- fect of high level acoustic excitation, grazing flow and bias flow have been suggested. One example is the model presented in [8] where the normalised impedance of a per- forate is expressed as

ZR = Re( ik σCc

[ lw

F(µ)+ δre

F(µ)fint]) + 1

σ[1 −2J1(kd) kd ] +(1− σ2

σ2Cc2 )1

2cVˆ+0.5

σ Mg+1.15

σCcMb (1)

ZI = Im( ik σCc

[ lw

F(µ)+ 0.5d

F(µ)fint]) − (1− σ2 σ2C2c ) 1

2c Vˆ

3

−0.3

σ Mg (2)

where ZR is the normalized resistance and ZI is the or- malized reactance, k is the wave number,σ is the poros- ity (percentage open area), Ccis the discharge coefficient,

lw is the plate thickness,µ is the adiabatic dynamic vis- cosity, µ = 2.179µ is the dynamic viscosity close to a conducting wall,ν= µ/ρ0is the kinematic viscosity, J is the Bessel function, d is the hole diameter, c is the speed of sound, Mg is the mean flow Mach number grazing to the liner surface, Mb is the bias flow Mach number in- side the holes of the perforate and ˆV is the peak value of the acoustic particle velocity in the hole. The rest of the parameters are defined as

K= r

−iω

ν (3)

F(µ) = 1 − 4J1(Kd/2)

Kd· J0(Kd/2) (4) δre= 0.2d + 200d2+ 16000d3 (5)

fint = 1 − 1.47√

σ+ 0.47√

σ3 (6)

Using Eqn. (1) and Eqn. (2) the magnitude of the terms related to high level nonlinear effects and bias flow, as well as grazing flow, can be compared. It should be noted that these terms are based on studies of the effect of nonlinearity and flow separately and not simultaneously.

The Cummings Equation

Consider orifice with bias flow, one of the most im- portant models to study the acoustic properties is Cum- mings [6] empirical equation. It is based on Bernoulli equation for unsteady flow, which in [25] is written as

l(t)dV dt + 1

2Cc2(U +V )|U +V | =P0+ d ˆPeiωt ρ0

(7)

where P0is the steady pressure drop; d ˆPeiωt is the sound pressure difference over the orifice; l(t)is variable effec- tive length of the fluid plug in the hole; V is oscillating velocity averaged over the plane of the orifice; U is the mean bias flow velocity in the orifice. For irrotaional flow,the length l(t) is 2l0+ lw, where l0≈ (π/8)d is the end-correction on one side of the orifice plane.When a jet is form,it according to Cummings [6] becomes a function of the effective length L:

Ł(τ) = Z τ

0 |U +V |dt (8)

10th International Conference on Flow-Induced Vibration (& Flow-Induced Noise)

(4)

where the time τ is measured from the prevous flow

“changeover” during the sign of U+V (t) is constant. The empirical coefficientε= l(t)/L0, where L0= 0.425d + lw

in [6] is the maximum possible lost end correction, were expressed as

ε= 1 − [1 + (L/d)1.585/3]−1 (9)

This function is a little different in [25] when con- sidering cases with bias flow. In the paper the effective length is divided into two parts: where the end correc- tion on the inflow side remainss constant as l0but on the outflow side decreases accordding to the jet length, as fol- lows

l(t) = l0+ l0+ lw

1+ (L/d)1.585/3 (10)

Thus, in the cases of U ≫ ˆV , Cummingss equation can be expressed as

l(t)dV dt + V

Cc2(U +V

2) = d ˆPeiωt ρ0

(11)

here l(t) → l0 for the jet lenth is large enough. At high frequency(ωR/U > 1),l(t) revert to the value 2l0+ lw,for unsteady volume flux through the aperture causes pulsa- tions in the jet cross-sectional area in or just downstream of the aperture.Using an effective hole thickness(l) which can vary between l0 and 2l0+ lw we get a normalized impedance

Z= d ˆP

ρ0c ˆV = ikl + U cCc2+ Vˆ

2cC2c (12)

This gives a Rayleigh conductivity

KR=ikπR2 Z =πR2

l

ωl/U (ωl/U) + i

C2c(1 + Vˆ 2U)

(13)

If this is linearlized a Rayleigh conductivity model was in [25] developed as

K= K0(ωl/U)

(ωl/U) + i/Cc2 (14)

where K0= πR2/l, l = 2l0+ lw

EXPERIMENTAL SETUP

FIGURE 1: Experimental configuration

The experimental configuration is illustrated in Fig. (1). The test object was an orifice plate with 3 mm thickness and 6 mm hole diameter. The orifice plate was mounted in a rigid tube with a diameter of 40 mm. On the left hand side, a high quality loudspeaker was mounted as the excitation source. Pure tone excitation was used and it was checked make sure that nonlinear harmonics gener- ated at the loudspeaker were sufficiently small. Two mi- crophones (1-4) were mounted on each side of the sample so that we could use two-microphone wave decomposi- tion to identify the sound wave components on each side.

In order to get the mean flow velocity passing through the orifice, two steady pressure sensors 1 2 were mounted to measure the steady pressure drop ∆P. The calculation is according to ISO5167-1:2003 [27], as fol- lows

U= Cc

s

2∆P

ρ0(1 − β4) (15)

whereβ = d/D. According to [27] the discharge coeffi- cient Ccshould be a function ofβ , Reynolds number and etc.. Here we suppose it is 0.75 recommended by Cum- mings [6]. In the study we consider two mean flow cases:

one is with ∆P = 14Pa and the mean flow velocity in the orifice is 3.65 m/s; the other is ∆P = 40Pa and the mean flow velocity is 6.17 m/s.

(5)

Under the plane wave assumption, sound wave com- ponents on both sides the orifice can be expressed as

 Pu+

Pu



=eik+d1 e−ikd1 eik+d2 e−ikd2

−1

 P1 P2



(16)

 Pd+

Pd



=eikd2 e−ik+d2 eikd1 e−ik+d1

−1

 P3 P4



(17)

where k= ω/c, M = U/c, k+= k/(1 + M), k= k/(1 − M). With acoustic waves amplitudes Pu+, Pu, Pd+, Pd, the oscillating velocity in the orifice ˆV and acoustic prop- erties, such as normalized impedance Z, Rayleigh con- ductivity KRcan be given as

Vˆ =Pu+− Pu

ρ0cσ (18)

Z= Pu++ Pu− Pd+− Pd

ρ0c ˆV (19)

KR= 2R ·ik· πR/2

Z (20)

RESULTS AND DISCUSSION Effects of in incident pressure level

FIGURE 2: Acoustic particle velocity in the orifice as a function of pressure difference level without bias flow

Fig. (2) shows the acoustic particle velocity for dif- ferent levels of pressure difference over the orifice. The

FIGURE 3: Normalized impedance in the orifice as a function of inverse Strouhal number ˆV/ωd

FIGURE 4: Rayleigh conductivity in the orifice as a func- tion of inverse Strouhal number ˆV/ωd

nonlinear behavior at higher levels of excitation can be clearly seen. The corresponding normalized impedance is shown in Fig. (3), and we can see the non-linearity at fairly low acoustic excitation levels. Instead plotting the Rayleigh conductivity makes the curves for different fre- quencies collapse, as can be seen in Fig. (4). The real part approaches the value K0/2R = 0.61 at low levels. The real part decrease when the inverse Strouhal number is near 1 and goes to a very low value at high inverse Strouhal num- bers; while the imaginary part first increases and reaches a maximum at an inverse Strouhal number around 1.5.

10th International Conference on Flow-Induced Vibration (& Flow-Induced Noise)

(6)

FIGURE 5: Normalised hole impedance divided by the Helmholtz number plotted against inverse Strouhal num- ber ˆV/ωd

It can be seen from Fig. (5) that dividing the nor- malized impedance by the Helmholtz number makes the curves for different frequencies collapse which is consis- tent with the result for the Rayleigh conductivity. There is also a fairly good agreement between experimental re- sistance and the results from both the Elnady and Cum- mings models. These two models for the resistance only differs on two points: The Cummings model does not in- clude any linear resistance term and the slope of the non- linear term is reduced by a factor 1− σ2 in the Elnady model compared to the Cummings model. The slope can in both models be changed by choosing another value for the discharge coefficient. Here the discharge coefficient has been set to 0.75 for both models. For the reactance term there is a reasonable agreement between experimen- tal results and the Elnady model at lower inverse Strohal numbers. Higher up the model does not catch the fact that the reactance does not continue to decrease with the increase in particle velocity. The Cummings model, as it is presented in Eqn. (12) does not include any nonlin- ear effect on the reactance term. In the article by Cum- mings [6] it is however discussed that the reactance may vary with time and with the effective jet length caused by the high level acoustic excitation. The effective thickness would at low acoustic levels take the value l= lw+2l0and would then decrease at higher acoustic excitation levels.

It can be seen that this agrees fairly well with the results in Fig. (5), the experimental reactance results start close to the curve for effective thickness l= lw+ 2l0 and ap- proaches the curve for effective thickness l= lw at high acoustic excitation levels.

Effects of bias flow velocity

FIGURE 6: Measured normalized impedance in the ori- fice as a function of ratio between a acoustic particle ve- locity and mean flow velocity

When there is a flow through the orifice the acoustic properties become more complex. Fig. (6) shows the nor- malized impedance as a function of the ratio of oscillating velocity to flow velocity. We divide the results into three parts according to the value of the velocity ratio: much less than unity (I), near unity (II) and much larger than unity (III). The resistance reduces as the velocity ratio in- creases in region I, has a minimum in region II and then increases in region II where the acoustic particle velocity dominates the behavior. The reactance has a more com- plex behavior and can either initially increase or decrease with increasing velocity ratio in region I. It then has a minimum in region II and the increases in region III to finally approach a constant value at high velocity ratios.

Fig. (7) compares the measured result with Elnady and Cummings models calculated by Eqn. (12). Neither resistance nor reactance is consistent with experimental data, especially near the ratio of unity. It seems the mach- anism is much more complex and further works should focus on the interaction of frequency, acoustic particle ve- locities and mean flow velocity.

In the conclusions of [25] it was mentioned that the Rayleigh conductivity for the case without flow reversal ( ˆV ≪ U) should approach the result in Eqn. (14). In or- der to check this experimental results for the Rayleigh

(7)

FIGURE 7: Normalized impedance in the orifice as a function of ratio between a acoustic particle velocity and mean flow velocity (U =6.17 m/s)

FIGURE 8: Rayleigh conductivity plotted against flow Strouhal numberωR/U

conductivity has been compared to the model results.

Fig. (8) shows the Rayleigh conductivity plotted against flow Strouhal number (ωR/U ). This means that at each Strouhal number there are a number of experimental data points representing different acoustic particle veloc- ity levels. It can be seen that the Rayleigh conductivity

does not exhibit a linear behavior since the results vary with acoustic excitation level at each flow Strouhal num- ber point. The agreement with the model result is also not very good. It can be seen that by varying the effec- tive hole thickness results of the right order of magnitude can be obtained but it seems that the Rayleigh conduc- tivity has a more complicated dependence on both mean flow velocity and acoustic excitation level than indicated by Eqn. (14).

CONCLUSIONS

An experimentally study of the acoustic properties for an orifice plate under high acoustic excitation levels and bias flow conditions has been made. Comparisons have been made with a semi-empiric impedance model [7, 8]

and the Cummings model as described in [25]. It was seen that without bias flow there is a reasonably good agreement between model results and measurements for the resistance. For the reactance the model according to [7, 8] catches the initial decrease with increasing exci- tation level but not the subsequent behavior at high excita- tion levels. The Cummings [6] model as described in [25]

discusses the possibility of an end correction which varies with both bias flow and high level acoustic excitation.

It can be seen that the measured reactance is within the range predicted by the suggested variations in end cor- rections. For the case with bias flow three regions were identified in terms of the ratio between acoustic particle velocity and mean flow velocity being: (I) smaller than unity, (II) around unity and (III) larger than unity. For re- gion I there was a decrease in resistance and a variation in reactance with velocity ratio. In region II both parts of the impedance had a minimum. In region III resistance increases while the reactance first has an increase and the approaches a constant value. Compared with experimen- tal data, it seems neither the Elnady nor the Cummings model gives a good prediction result since the nonliear acoustic machanism with bias flow is much more com- lex than that without. In [25] it was predicted that the Rayleigh conductivity would go to the linearized value according to Eqn. (14) for cases when the acoustic parti- cle velocity is smaller than the mean flow velocity in the orifice so that no flow reversal occurs. Comparisons with experimental results shows that this is not the case there is still a nonlinear variation in Rayleigh conductivity even when the velocity ratio is small.

REFERENCES

[1] Sivian, L. J., 1935. “Acoustic impedance of small orifices”. Journal of the Acoustical Society of Amer-

10th International Conference on Flow-Induced Vibration (& Flow-Induced Noise)

(8)

ica, 7, pp. 94–101.

[2] Ingrd, U., and Labate, S., 1950. “Acoustic circu- lation effects and the nonlinear impedance of ori- fices”. Journal of the Acoustical Society of America, 22, pp. 211–219.

[3] Ingrd, U., and Ising, H., 1967. “Acoustic nonlinear- ity of an orifice”. Journal of the Acoustical Society of America, 42, pp. 6–17.

[4] Melling, T., 1973. “The acoustic impendance of per- forates at medium and high sound pressure levels”.

Journal of Sound and Vibration, 29(1), pp. 1 – 65.

[5] Zinn, B., 1970. “A theoretical study of non-linear damping by helmholtz resonators”. Journal of Sound and Vibration, 13(3), pp. 347 – 356.

[6] Cummings, A., 1986. “Transient and multiple frequency sound transmission through perforated plates at high amplitude”. Journal of the Acousti- cal Society of America, 79, pp. 942–951.

[7] Elnady, T., and Boden, H., 2003. “On semi- empirical liner impedance modeling with grazing flow”. AIAA, pp. 1–11.

[8] Elnady, T., 2004. “Modelling and characterization of perforates in lined ducts and mufflers”. PhD The- sis, Department of Aeronautical and Vehicle Engi- neering, KTH, Stockholm, Sweden.

[9] Ingrd, U., 1970. “Nonlinear distortion of sound transmitted through an orifice”. The Journal of the Acoustical Society of America, 48(1A), pp. 32–33.

[10] TAM, C., and KURBATSKII, K., 2000. “Microfluid dynamics and acoustics of resonant liners”. AIAA Journal, 38, pp. 1331–1339.

[11] C.K.W. TAM, K.A. KURBATSKII, K. A., and JR, R. G., 2001. “A numerical and experimental inves- tigation of the dissipation mechanisms of resonant acoustic liners”. Journal of Sound and Vibration, 245(3), pp. 545 – 557.

[12] D. Bechert, U. M., and Pfizenmarier, E., 1977.

“Experiments on the transmission of sound through jets”. In American Institute of Aeronautics and As- tronautics, Aeroacoustics Conference, 4th, Atlanta.

[13] Howe, M., 1979. “On the theory of unsteady high reynolds number flow through a circular aper- ture”. Royal Society of London Proceedings A, 366, pp. 205–233.

[14] Howe, M., 1980. “The dissipation of sound at an edge”. Journal of Sound and Vibration, 70, pp. 407–

411.

[15] Howe, M., 1998. Acoustics of Fluid-structure inter- actions. Cambridge University Press.

[16] Hughes, I. J., and Dowling, A. P., 1990. “The ab- sorption of sound by perforated linings”. Journal of

Fluid Mechanics, 218, pp. 299–335.

[17] Eldredge, J. D., and Dowling, A., 2003. “The ab- sorption of axial acoustic waves by a perforated liner with bias flow”. Journal of Fluid Mechanics, 485, pp. 307–335.

[18] Ronneberger, D., 1972. “The acoustical impedance of holes in the wall of flow ducts”. Journal of Sound and Vibration, 24, pp. 133–150.

[19] Ronneberger, D., 1980. “The dynamics of shearing flow over a cavity a visual studyrelated to the acous- tic impedance of small orifices”. Journal of Sound and Vibration, 71, pp. 565–581.

[20] Walker, B., and Charwat, A., 1982. “Correlation of the effects of grazing flow on the impedance of helmholtz resonators”. Journal of the Acoustical So- ciety of America, 72, pp. 550–555.

[21] Nelson, N. H., and Doak, P., 1981. “Fluid dynam- ics of a flow excited resonance, part i: Experiment”.

Journal of Sound and Vibration, 78, pp. 15–38.

[22] Nelson, N. H., and Doak, P., 1983. “Fluid dynamics of a flow excited resonance, part ii: Flow acoustic interaction”. Journal of Sound and Vibration, 91, pp. 375–402.

[23] Worraker, W., and Halliwell, N., 1985. “Jet engine liner impedance: An experimental investigation of cavity neck flow/acoustics in the presence of a mach 0.5 tangentialshear flow”. Journal of Sound and Vi- bration, 103, pp. 573–592.

[24] Goldman, A., and Panton, R., 1976. “Measurement of the acoustic impedance of an orifice under a tur- bulent boundary layer”. Journal of the Acoustical Society of America, 60, pp. 1397–1404.

[25] T. Luong, M. H., and McGowan, R., 2005. “On the rayleigh conductivity of a bias-flow aperture”.

Journal of Fluids and Structures, 21(8), pp. 769 – 778.

[26] J. Rupp, J. C., and Spencer, A., 2010. “Interaction between the acoustic pressure fluctuations and the unsteady flow field through circular holes”. Journal of Engineering for Gas Turbines and Power, 132(6), p. 061501.

[27] Iso5167-1:2003 measurement of fluid flow by means of pressure differential devices inserted in circular cross-section conduits running full.

References

Related documents

the loudspeaker voltage and amplitude of incident pressure wave (p + ), at 110 Hz remained unchanged. Figures 9 and 10 shows a subset of the two tone excitation results at 110

Figure 10 shows normalized impedance results obtained using multi-tone excitation on sample P4 with 2 % porosity, plotted against total peak particle velocity in the

In the linear case the impedance is independent of the sound field but when the sound pressure level is high the perforate impedance will be dependent on the acoustic

(3) gives an under prediction of the drop in effective orifice length caused by the high level acoustic excitation. The model can give reasonable results when compared to

The bias factor of an edge dislocation can be calculated by solving the dif- fusion equation with a drift term if the interaction energy between the dislocation and a mobile defect

We find that: (i) although principles of need and cost-effectiveness may recommend the same allocation of resources the underlying reason for an allocation is different; (ii) while

Trots negativ attityd och inställning innan implementering ( Brooker, Peters, McCabe & Short, 1999 ; Indig, Copeland, Conigrave & Rotenko, 2009 ; Karlsson,

So the result of the study should show that the US British and Swedish media takes on a negative tone towards Hugo Chavez and that the Chinese and Russian media instead takes on