12
The pulse tube and the pendulum G. W. Swift and S. Backhaus Condensed Matter and Thermal Physics Group, Los Alamos National Laboratory, Los Alamos, New Mexico 87545 Received 2 June 2009; revised 17 August 2009; accepted 20 August 2009 An inverted pulse tube in which gravity-driven convection is suppressed by acoustic oscillations is analogous to an inverted pendulum that is stabilized by high-frequency vibration of its pivot point. Gravity acts on the gas density gradient arising from the end-to-end temperature gradient in the pulse tube, exerting a force proportional to that density gradient, tending to cause convection when the pulse tube is inverted. Meanwhile, a nonlinear effect exerts an opposing force proportional to the square of any part of the density gradient that is not parallel to the oscillation direction. Experiments show that convection is suppressed when the pulse-tube convection number N ptc = 2 a 2 T / T avg / gD sin - L cos is greater than 1 in slender tubes, where is the radian frequency of the oscillations, a is their amplitude, T is the end-to-end temperature difference, T avg is the average absolute temperature, g is the acceleration of gravity, L is the length of the pulse tube and D is its diameter, is about 1.5, and the tip angle ranges from 90° for a horizontal tube to 180° for an inverted tube. Theory suggests that the temperature dependence should be T / T avg instead of T / T avg 2009 Acoustical Society of America. DOI: 10.1121/1.3238156 PACS numbers: 43.35.Ud, 43.25.Ts, 43.28.Py RR Pages: 2273–2284 I. INTRODUCTION Rigid pendula exhibit many interesting phenomena, 1 in- cluding dynamic stabilization: If the pivot point of a rigid pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi- bration axis, and the pendulum can stand inverted, seeming to defy gravity. The equation of motion can be derived with the simple approach of Blitzer. 2 Let the x and y directions, the angle of the pivot-point vibration relative to gravity g, and the angle t of the pendulum relative to gravity be defined as shown in Fig. 1a, and let m be the mass of the pendulum’s bob and L be the length of its rod. The positions xt and yt of the pendulum’s bob are given by x = a sin cos t + L sin , 1 y =- a cos cos t - L cos , 2 when the pivot point is forced to vibrate sinusoidally with amplitude a and radian frequency . The equation of motion for the bob can be written as F x = mx ¨ , F y - mg = my ¨ , 3 where the F’s are the two components of force exerted by the rod on the bob and the overdots represent time derivatives. By Newton’s third law, the bob exerts force -F on the rod, applied at the end of the rod attached to the bob. The torque on the rod about the pivot point due to -F must be zero, because this massless rod has zero moment of inertia, I rod . Thus - LF x cos - LF y sin = I rod ¨ = 0. 4 Combining these four equations by eliminating x, y, F x , and F y yields the equation of motion for the pendulum angle as follows: ¨ =- g/Lsin + 2 a/Lsin - cos t . 5 For 0 180°, the torque applied by gravity that tends to decrease is apparent in Fig. 1a and Eq. 5. The time-averaged torque caused by the pivot-point vibration, tending to align with the vibration, is not so apparent in Fig. 1a or Eq. 5, but Fig. 1b helps explain the mecha- nism, if the vibration is exaggerated and gravity is neglected. 1 The figure shows the pendulum at two extremes j and k of its motion under this exaggerated circumstance. At j , the acceleration of the pivot point causes a large, positive ˙ , without much acceleration of the bob. At k, the accelera- tion of the pivot point causes the bob to accelerate parallel to the vibration direction, with only a small, negative ˙ . The net effect on is positive, causing the pendulum to tend to align with the vibration. In other words, the time-averaged torque tending to align with the vibration is proportional to the product of the amplitude of the angular vibration, k - j , and how strongly the torque caused by the vibration varies with . For decades, quantitative analysis of Eq. 5 when 2 g / L has appeared as an exercise in textbooks on classical mechanics, such as Ref. 3 for = 180° and =90° and Ref. 4 for = 180°. Extended to arbitrary , the analysis shows that the dimensionless number N pendulum = 2 a 2 gL 6 determines the simple pendulum’s behavior for slow time scales, i.e., time scales 1 / . For N pendulum 4, the pendu- lum can be stably held up for any . For 1 N pendulum 4, it can be stably held up only if the pivot-point’s vibration is close enough to vertical i.e., close enough to either =0° or = 180°, which are equivalent. For N pendulum 1, the pendu- lum cannot be stably inverted for any . The details of this J. Acoust. Soc. Am. 126 5, November 2009 © 2009 Acoustical Society of America 2273 0001-4966/2009/1265/2273/12/$25.00

The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

The pulse tube and the pendulumG. W. Swift and S. BackhausCondensed Matter and Thermal Physics Group, Los Alamos National Laboratory, Los Alamos,New Mexico 87545

�Received 2 June 2009; revised 17 August 2009; accepted 20 August 2009�

An inverted pulse tube in which gravity-driven convection is suppressed by acoustic oscillations isanalogous to an inverted pendulum that is stabilized by high-frequency vibration of its pivot point.Gravity acts on the gas density gradient arising from the end-to-end temperature gradient in thepulse tube, exerting a force proportional to that density gradient, tending to cause convection whenthe pulse tube is inverted. Meanwhile, a nonlinear effect exerts an opposing force proportional to thesquare of any part of the density gradient that is not parallel to the oscillation direction. Experimentsshow that convection is suppressed when the pulse-tube convection number Nptc

=�2a2��T /Tavg / �g��D sin �−L cos ��� is greater than 1 in slender tubes, where � is the radianfrequency of the oscillations, a is their amplitude, �T is the end-to-end temperature difference, Tavg

is the average absolute temperature, g is the acceleration of gravity, L is the length of the pulse tubeand D is its diameter, � is about 1.5, and the tip angle � ranges from 90° for a horizontal tube to 180°for an inverted tube. Theory suggests that the temperature dependence should be �T /Tavg instead of��T /Tavg. © 2009 Acoustical Society of America. �DOI: 10.1121/1.3238156�

PACS number�s�: 43.35.Ud, 43.25.Ts, 43.28.Py �RR� Pages: 2273–2284

I. INTRODUCTION

Rigid pendula exhibit many interesting phenomena,1 in-cluding dynamic stabilization: If the pivot point of a rigidpendulum is vibrated at high enough frequency and highenough amplitude, the pendulum tends to align with the vi-bration axis, and the pendulum can stand inverted, seemingto defy gravity. The equation of motion can be derived withthe simple approach of Blitzer.2 Let the x and y directions,the angle � of the pivot-point vibration relative to gravity g,and the angle ��t� of the pendulum relative to gravity bedefined as shown in Fig. 1�a�, and let m be the mass of thependulum’s bob and L be the length of its rod. The positionsx�t� and y�t� of the pendulum’s bob are given by

x = a sin � cos �t + L sin � , �1�

y = − a cos � cos �t − L cos � , �2�

when the pivot point is forced to vibrate sinusoidally withamplitude a and radian frequency �. The equation of motionfor the bob can be written as

Fx = mx, Fy − mg = my , �3�

where the F’s are the two components of force exerted by therod on the bob and the overdots represent time derivatives.By Newton’s third law, the bob exerts force −F� on the rod,applied at the end of the rod attached to the bob. The torqueon the rod about the pivot point due to −F� must be zero,because this massless rod has zero moment of inertia, Irod.Thus

− LFx cos � − LFy sin � = Irod� = 0. �4�

Combining these four equations by eliminating x, y, Fx, andFy yields the equation of motion for the pendulum angle � as

follows:

J. Acoust. Soc. Am. 126 �5�, November 2009 0001-4966/2009/126�5

� = − �g/L�sin� + �2�a/L�sin �� − ��cos �t . �5�

For 0���180°, the torque applied by gravity thattends to decrease � is apparent in Fig. 1�a� and Eq. �5�. Thetime-averaged torque caused by the pivot-point vibration,tending to align � with the vibration, is not so apparent inFig. 1�a� or Eq. �5�, but Fig. 1�b� helps explain the mecha-nism, if the vibration is exaggerated and gravity isneglected.1 The figure shows the pendulum at two extremes jand k of its motion under this exaggerated circumstance. Atj, the acceleration of the pivot point causes a large, positive�, without much acceleration of the bob. At k, the accelera-tion of the pivot point causes the bob to accelerate parallel tothe vibration direction, with only a small, negative �. Thenet effect on � is positive, causing the pendulum to tend toalign with the vibration. In other words, the time-averagedtorque tending to align � with the vibration is proportional tothe product of the amplitude of the angular vibration, �k

−� j, and how strongly the torque caused by the vibrationvaries with �.

For decades, quantitative analysis of Eq. �5� when �2

�g /L has appeared as an exercise in textbooks on classicalmechanics, such as Ref. 3 for �=180° and �=90° and Ref. 4for �=180°. Extended to arbitrary �, the analysis shows thatthe dimensionless number

Npendulum =�2a2

gL�6�

determines the simple pendulum’s behavior for slow timescales, i.e., time scales �1 /�. For Npendulum�4, the pendu-lum can be stably held up for any �. For 1�Npendulum�4, itcan be stably held up only if the pivot-point’s vibration isclose enough to vertical �i.e., close enough to either �=0° or�=180°, which are equivalent�. For Npendulum�1, the pendu-

lum cannot be stably inverted for any �. �The details of this

© 2009 Acoustical Society of America 2273�/2273/12/$25.00

Page 2: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

analysis are omitted here, because similar details are pre-sented in the next paragraph and in Sec. II.�

A more complicated pendulum sets the stage well foranalysis of the pulse tube in Sec. II. Figure 1�c� shows a ringof radius L and mass m, supported from its central pivotpoint by massless spokes. The mass m is uniformly distrib-uted around the ring, except for an out-of-balance part�m /2, which has been removed from location −� and addedto location +�. The pivot point is forced to vibrate sinusoi-dally along a line at angle � from the vertical, with amplitudea and angular frequency �, as for the simple pendulum con-sidered above, and the analysis again begins by followingBlitzer’s approach.2 The coordinates of the “positive” mass

FIG. 1. �a� An inverted, rigid pendulum whose support is vibrated at a highenough frequency and amplitude can be stabilized against falling over. Rela-tive to gravity, which points in the −y direction, � gives the angle alongwhich the vibrations occur, and ��t� gives the angle of the pendulum. �b�Consideration of an exaggerated situation, in which the amplitude of thepivot-point oscillation is not small compared with the length of the pendu-lum’s rod, illustrates the stabilization mechanism. �c� An out-of-balance ringthat is vibrated at its central pivot point behaves similarly, due to the samemechanism, but described by slightly more complicated mathematics. �d� Aninverted pulse tube whose gas is oscillated at a high enough frequency andamplitude can be stabilized against natural convection, because of a similarmechanism. The double-headed arrows show the peak-to-peak amplitude 2aof the acoustic oscillation, assumed to carry all isotherms. Ideally, the cen-tral slug of gas, whose length is L−2a, would experience no gravity-drivenconvection.

at +� and the “negative” mass at −� are written down, as are

2274 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

equations of motion for these two masses. The torque exertedon the ring by its contact with the two out-of-balance massesis set equal to Iring�, where Iring=mL2 is the ring’s moment ofinertia. Eliminating the forces and coordinates yields anequation of motion for the angle ��t�,

mL2� = − �mgL sin� + �m�2aL sin �� − ��cos �t �7�

�grav + vib, �8�

where the right-hand side can be interpreted as the sum of atorque due to gravity and a torque due to the vibration of thepivot. References 3 and 4 show that the pendulum angle �should be written as the sum of slow and fast parts

� = �slow + �fast cos �t , �9�

when �2��mg /mL. Using Taylor-series expansions

= ��slow+ � �

���

�slow

�fast cos �t �10�

for both terms on the right-hand side of Eq. �7� and timeaveraging confirms that there was no need for a sin �t termin Eq. �9� and it generates two equations: a “fast” equationfrom the cos �t terms and a “slow” equation from terms withnonzero time-averages,

�fast = −�m

m

a

Lsin�� − �slow� , �11�

mL2�slow = − �mgL sin �slow

+��a�m�2

4msin�2� − 2�slow� . �12�

If this pendulum can be stably inverted, then �slow=0 andsetting the right-hand side of Eq. �12� equal to zero mustyield a solution for �slow. Whether such a solution exists, andits value, depends on the ratio of the coefficients of the twoterms on the right �dropping the factor of 4 for simplicity�,

Nring =�2a2

gL

�m

m. �13�

It is interesting how many effects combine to put �m and��m�2 /m in the two terms on the right-hand side of Eq. �12�,and hence to put �m /m in Eq. �13�. The left-hand side of Eq.�12� is the moment of inertia times the slow angular accel-eration, so the right-hand side must be the slow torque. Thefirst term on the right-hand side shows gravity applyingtorque proportional to the unbalanced mass �m, with themass �m /2 trying to drop down on the right and the negativemass −�m /2 trying to float up on the left, both applyingclockwise torques in Fig. 1�c�. The second term on the right-hand side, which represents the time-averaged torque tendingto align the unbalanced axis of the ring with �, arises fromthe mechanism illustrated in Fig. 1�b�, i.e., the time-averagedproduct of �vib /�� and �fast in Eq. �10�, where vib is thevibration torque given by the second term in Eq. �7�. Theunbalanced mass �m appears in the numerator of �fast in Eq.�11� because it is responsible for vib through the mechanism

shown in Fig. 1�b�. The ring mass m appears in the denomi-

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum

Page 3: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

nator of �fast in Eq. �11� because the ring’s moment of inertiaresists the fast torque. Finally, �vib /�� retains the mass de-pendence of vib itself, namely, �m. Overall, Nring representsthe angle-independent part of the ratio of the time-averagedalignment torque �proportional to ��m�2 /m� to the gravita-tional torque �proportional to �m�.

This paper explores the hypothesis that a similar mecha-nism is responsible for suppressing natural convection in thegas in a tube with an axial temperature gradient, when thegas oscillates axially at high enough frequency and ampli-tude. As shown in Fig. 1�d� for such a tube with its cold endhigher than its hot end, gravity tends to pull the dense gasnear the cold end to one side and down, pushing the less-dense gas near the hot end to the other side and up. This putsthe center of mass of the gas below the tube’s centerline. Thevibration can then exert a time-averaged torque on the entirebody of gas, via time-averaged oscillating forces on this off-center center of mass, in a process analogous to that shownin Fig. 1�b�. This torque opposes that of gravity and canbalance it, preventing convection.

The gas in a pulse tube experiences such axial oscilla-tions and supports such an axial temperature gradient. Thepulse tube, a vital component of cryogenic pulse-tuberefrigerators,5 is a smooth-walled tube without internal struc-tures, bounded on both ends by flow straighteners and heatexchangers through which gas flows easily. Its purpose is totransmit acoustic power through the gas from the cold end tothe hot end with minimal heat leak from the hot end to thecold end. The lack of internal structure generally makes low-loss transmission of acoustic power easy, but makes heat-leak minimization challenging. The peak-to-peak volumetricstroke of the moving gas is always less than the volume ofthe gas in a pulse tube. Ideally, one imagines a perfectlythermally stratified slug of gas, whose volume is the differ-ence between the tube volume and the volumetric stroke,oscillating axially in the tube and remaining entirely insidethe tube at all times, conducting only a little heat from hot tocold, without any accompanying convection. However, sev-eral heat-transfer mechanisms can disturb this ideal picture,carrying much more heat than would be carried by conduc-tion alone, unless attention is paid to minimizing each one ofthem. One such heat-transfer mechanism—natural convec-tion due to gravity acting on density gradients in the gas, thesubject of this paper—is known to occur commonly in low-frequency pulse-tube refrigerators, but it is also known thatsuch convection is often reduced or absent in high-frequencypulse-tube refrigerators.6–8

Our motivation for this work arose in the context ofpulse-tube refrigerators and thermoacoustic engines, some-times coupled, in which convectively stable orientation ofthe tubes relative to gravity was inconvenient and an accu-rate understanding of the suppression of convection by high-frequency oscillations was desired. In thermoacoustic-Stirling hybrid engines9 and cascade thermoacousticengines,10 the tubes that transmit acoustic power across atemperature difference while minimizing heat leak are calledthermal-buffer tubes. They generally carry acoustic powerfrom a hot temperature to ambient temperature, while pulse

tubes carry acoustic power from a cold temperature to ambi-

J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009 G. W.

ent temperature. But even in pulse-tube refrigerators, thesetubes are sometimes called thermal-buffer tubes. For brevityin this paper, all such tubes are referred to as pulse tubes, andtheir end temperatures are labeled as hot and cold.

Below, theoretical arguments �Sec. II� and experimentalevidence �Sec. III� are presented to show that

Nptc =�2a2

g��D sin � − L cos �� �T

Tavg

�14�

is a useful and plausible choice of dimensionless group forcharacterizing this phenomenon in pulse tubes of low-aspectratio D /L. As above, � is the radian frequency of the oscil-lation, a is its displacement amplitude, and g is the accelera-tion of gravity; D is the pulse-tube diameter and L is itslength, �T is the end-to-end temperature difference, and Tavg

is the average temperature. The tip angle � is taken to be zeroin the vertical, gravitationally stable orientation, and thisequation is only valid for 90° ���180° �where cos ��0,so both terms in the denominator are non-negative�. The pa-rameter � is a fitting parameter discussed below, experimen-tally found to be about 1.5, and experiment shows that isclose to 1/2 while theory suggests =1.

II. THEORY

An extensive literature describes the interaction betweenrapid vibration and steady convection in fluids, in the frame-work of the Boussinesq approximation, namely, that densityvariations due to temperature variations are small and den-sity variations due to pressure variations are zero. This lit-erature is reviewed and its foundations are succinctly sum-marized by Gershuni and Lyubimov.11 After writing thehydrodynamic and thermal variables as the sum of fast varia-tions at the vibration frequency and slow variations, theyderive time-averaged equations of motion for the slow vari-ables similar in spirit to Eq. �12� above, showing that fastvibrations effectively add a time-averaged body force to thefluid, whose magnitude and direction depend on the magni-tude and direction of the vibration velocity and the fluid’stemperature-gradient vector field. For steady state with neg-ligible convection, their Eqs. �1.100� and �1.101� give theconditions for balance between gravity- and vibration-induced forces

Ra�� T � g + Ravib�� �w� · n� � �� T = 0, �15�

�� · w� = 0, �16�

�� � w� = �� T � n , �17�

where g and n are unit vectors in the directions of gravityand the vibration, respectively, T is the time-averaged tem-perature, w� is the solenoidal part of Tn, and the ordinaryRayleigh number Ra and the vibrational Rayleigh numberRavib are given by

Ra =l3g�Tchar , �18�

�Tchar

Swift and S. N. Backhaus: The pulse tube and the pendulum 2275

Page 4: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

Ravib =��al�Tchar�2

2 �Tchar2 , �19�

with l a characteristic length of the boundary of the fluid,�Tchar a characteristic temperature difference, Tchar a charac-teristic temperature, a characteristic kinematic viscosity,and � a characteristic thermal diffusivity. In Eqs. �18� and�19�, we have set the thermal expansion coefficient of Ref.11 equal to 1 /T, as appropriate for an ideal gas. Evident fromEq. �15�, the existence of a steady state without convectiondepends on the ratio of Ravib and Ra,

�2a2

gl

�Tchar

Tchar. �20�

Although derived in the context of the Boussinesq approxi-mation, which is not really applicable to pulse tubes, thisexpression suggests most of the functional dependences thatare displayed in Eq. �14�, most of which are confirmed in theexperiments described below. Presumably, numerical analy-sis based on Eqs. �15�–�17� could show whether the pulse-tube’s length L, its diameter D, or some combination of thosevariables is best used for the characteristic length l, andcould find the tip-angle dependence of vibrational suppres-sion of convection in a pulse tube.

A high vibrational Rayleigh number tends to align den-sity gradients along the direction of vibration, whether or notgravity is involved. Thus, we expect that this phenomenonalso mitigates the effect of jet-driven streaming due to im-perfect flow straightening and the effect of Rayleigh stream-ing, on Earth in zero gravity, because both of these streamingphenomena create non-axial density gradients in pulsetubes.12 However, since streaming grows more intense as �arises, the mitigation cannot be as abrupt a function of �a asit is for gravity-driven convection. Nevertheless, at a given�T, the effect of streaming might be reduced significantly.

The rest of this section presents a very simple attempt toanticipate the best choice for l in Eq. �20� when the pulse-tube’s length L is significantly greater than its diameter D,which is a common situation in pulse-tube refrigerators. Al-though the approximations used here might seem crude, wehope that they can correctly capture the dominant functionaldependences on D and L.

Three characteristic times are well separated. For a typi-cal sinusoidally driven pulse-tube refrigerator, 1 /��0.003 s. This is significantly faster than the time requiredfor an appreciable change in convective motion, estimatedfrom the ring-pendulum analysis to be of the order of �l /g�0.1 s. This, in turn, is significantly faster than the diffusivethermal-relaxation time l2 /��30 s. Thus, for rough esti-mates, it is plausible to assume that temperatures are essen-tially carried with the moving gas on the time scales of thegas motion and that the dynamical behavior of the gravity-vibration interaction in the gas is qualitatively similar to thatof the ring pendulum.

Furthermore, since �� in gases, the viscous relaxationtime for l-scale distance is also �30 s, so the viscous pen-etration depth �2 /� is typically much smaller than l. Thevelocity of the developing steady flow might be of the order

of l / l /g, so the steady-flow Reynolds number might ini-

2276 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

tially be roughly �l2 / � /�l /g�300, a regime in whichinertial effects are important and two- and three-dimensionalflows are often time dependent. The typical Rayleigh numbergiven in Eq. �18� can be estimated as �l2 / ��l2 /�����Tchar /Tchar� / �l /g��106�Tchar /Tchar, so modest �Tchar /Tchar can cause significant convection. Similarly, the typicalvibrational Rayleigh number in Eq. �19� can be estimated as108�a / l�2��Tchar /Tchar�2, so values of a / l that are common inpulse tubes can make Ravib�Ra.

To keep the analysis of the problem simple, we retainthe Boussinesq approximation, treating the gas in the pulsetube as incompressible. Thus, the double-headed arrows inFig. 1�d�, illustrating the peak-to-peak stroke of the gas, aretaken to be the same length at the two ends of the pulse tube.The isotherms in Fig. 1�d� are shown at an instant of timewhen the motion of the gas is at mid-stroke, e.g., when �t=−� /2. A quarter cycle later, cos �t=1 and the uppermostisotherm would have just touched the cold heat exchanger;another half cycle later, when cos �t=−1, the lowermost iso-therm would just touch the hot heat exchanger. The slug ofgas between these two isotherms, which always remains in-side the pulse tube, is the object of interest. It has a lengthL−2a, which we might take to be the effective length for thisproblem. However, our experiments cannot resolve the smalldifference between this length and L itself, so for simplicitywe use L in the rest of this derivation.

The uppermost isotherm has temperature TC when it ismomentarily in contact with the cold heat exchanger at thattemperature, but the pressure-induced adiabatic heating andcooling that the gas experiences causes its average tempera-ture to be TC,avg=TC�1+ ��−1�pa sin /�pm�, where � is theratio of isobaric to isochoric specific heats, pa is the pressureamplitude, pm is the mean pressure, and =� /2 is the phaseby which oscillating pressure leads oscillating velocity �posi-tive velocity going from hot to cold�.13 The hot isotherm’stemperature TH,avg obeys a similar expression. Our experi-ments cannot resolve the effects of these small pa-dependenttemperature differences, so for simplicity we describe thetemperatures of the slug of gas with �T=TH−TC and Tavg

= �TH+TC� /2 in the rest of this derivation, instead of similarbut more complicated expressions with TH,avg and TC,avg.

As shown in Fig. 2�a�, imagine that motion within thisslug of gas in the pulse tube can be modeled as plug flow ina loop of piping that vibrates along the � direction and whosecross-sectional area is half of the cross-sectional area A ofthe pulse tube itself, so gas rising on the left half of the pulsetube in Fig. 1�d� is modeled as rising plug flow in the left legof Fig. 2�a�, and similarly down on the right. In this model,the convective motion in the pulse tube is represented by asingle degree of freedom, measured by a time-dependent dis-placement ��t�. This displacement and the superimposed vi-bration carry the isotherms, because the thermal-relaxationtime is so much slower than the times for these motions, asestimated above. Then 2� is the measure of how far theisotherms in the right half of the loop are misaligned fromthose in the left half at any instant of time, with the sign of �as shown in the figure. Ignoring end effects for small �, andassuming that end-to-end temperature differences are small

enough that the density � can be assumed to be essentially

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum

Page 5: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

linear in position �not obviously a good assumption, but lin-ear T�z� is discussed below�, the density in the two legs ofthe loop can be written as

��z� = �H + �z � ����/L �21�

except near the ends, where z is the distance from the hotend, ��=�C−�H, the plus sign is chosen for the right leg andthe minus sign for the left leg, and the subscripts on � cor-respond to those on T above. Thus, when �=0, the densityrises linearly from �H at z=0 to �C at z=L in both legs, andnonzero � shifts one of these density profiles up and the otherone down.

A Lagrangian derivation of the equation of motion for��t� is well suited to keeping track of details here. The ap-plied vibrational displacement a cos �t is superimposed onthe plug-flow displacement �, so the velocity of the gas is

−�a sin �t+ � in the left leg of the loop and −�a sin �t− �in the right leg. Transverse kinetic energy near the ends, andother end corrections to the kinetic energy, are neglected be-

FIG. 2. �a� The convective flow in the pulse tube can be modeled crudely asplug flow in a loop of pipe, characterized by a single degree of freedom,measured by �. As shown, the plug-flow displacement � creates a misalign-ment of 2� between isotherms on the left and right legs of the loop. Achange in � can be caused by gravity, or by the vibration acting on anoff-center center of mass caused by nonzero � itself. �b� The effect of achange in � on the gravitational potential energy of the gas in such a loop ofpipe can be estimated from an even more simplified model.

cause D�L is assumed. Then the total kinetic energy is

J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009 G. W.

K =1

2��avg − ���/L�

A

2L�− �a sin �t + ��2

+1

2��avg + ���/L�

A

2L�− �a sin �t − ��2 �22�

=AL

2�avg��2a2 sin2�t + �2� + A�����a sin �t . �23�

In Eq. �22�, the first term is the kinetic energy in the left legof the loop, and the second term that in the right leg. Thedensity factors in these terms come from averaging Eq. �21�with respect to z.

The potential energy change U due to � can be estimatedby considering Fig. 2�b�. As � changes from zero to a non-zero value, isotherms far from the ends of the tube contributeno change to U, because for any mass moving up in the leftleg there is an equal mass associated with the same isothermmoving down the same distance in the right leg. The samecancellation would occur for the gas within � of the end ofthe tube, if it did not have to “turn the corner,” changingfrom the left leg to the right leg at the top or the right to theleft at the bottom; if such gas parcels could move to thepositions shown as crosshatched in Fig. 2�b�, their effects onU would be canceled by their partners of the same isothermsin the other leg. Thus, the net effect of nonzero � is to lowersome cold gas whose mass is of the order of �C�A /2�� adistance of the order of D sin �−� cos � and raise some hotgas whose mass is of the order of �H�A /2�� a similar dis-tance, yielding

U � − ��A�

2g 4

3�D sin � − � cos � , �24�

where the 4 /3� comes from careful consideration of thesemicircular cross section of each leg.

With the standard Lagrangian methods of classical me-chanics, the equation of motion for � is obtained by writing

�d /dt�� �K−U� /��−��K−U� /��=0. Using Eqs. �23� and�24� above for K and U, the result is

�avgAL� =A��g

2 4

3�D sin � − 2� cos �

− A���2a� cos �t . �25�

This equation resembles Eq. �7� for the unbalanced-ring pen-dulum. The total mass �avgAL in the loop of piping acceler-ates in the � direction in response to forces of gravity, ex-pressed by the first term, and in response to forces caused byvibration, expressed by the second term. Following the sameprocedure as for the unbalanced-ring pendulum, this equa-tion of motion is broken down into fast and slow parts bysubstituting �=�slow+�fast cos �t and assuming �fast��slow

and �2���g /�avgL. The fast part of � is then

�fast =��

�avg

a

L�slow, �26�

and the slow response of � to gravity and to the time-averaged product of �fast and the imposed vibration is de-

scribed by

Swift and S. N. Backhaus: The pulse tube and the pendulum 2277

Page 6: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

�avgAL�slow =A��g

2 4

3�D sin � − 2�slow cos �

−A����2�2a2�slow

2�avgL. �27�

If the vibrations suppress convection, then �slow=0, and thephenomenon should be governed by the surviving terms onthe right-hand side. Solving for �slow yields

�slow =4D sin �/3�

�2a2��/�avggL + 2 cos �. �28�

Too large a value of �slow would be unrealistic, becauseit would put the off-center cold gas and the off-center hot gasin Figs. 1�d� and 2�b� close together, thermally short-circuiting the temperature difference responsible for thevibration-stabilization effect. Thus, a stable �slow can be nolarger than some fraction of L, which can be convenientlywritten as 2L /3��, where � is as yet unknown. Making thatsubstitution for �slow in Eq. �28�, rearranging, and defining adimensionless group of variables resembling Eq. �14� yield

Nptc ��2a2

g��D sin � − L cos ����

�avg= 2. �29�

Since �� /�avg=�T /Tavg, this supports the dependencesshown in Eq. �14� above, for =1. Note that this derivationis valid for 90° ���180°, so the geometrical factor in thedenominator could just as well be written as �D�sin ��+L�cos ��.

Equations �28� and �29� are only valid for �a largeenough to suppress convective motion. For very large �a,�slow is generally small, as illustrated in Fig. 2�b�. However,if �a is just below the threshold, �slow could be fairly largeand essentially time dependent, and the picture of Fig. 2�b�would be unrealistic because the off-center slugs of extreme-temperature gas would extend over appreciable lengths, andtheir temperatures would no longer be uniformly at TC andTH, but rather would be distributions of less-extreme tem-peratures determined by competing conduction to both heatexchangers and between the two legs of the loop. Whetherthis might lead to Nptc���T /Tavg�, where �1, is notclear. Further analysis of this issue may require numericalstudy of Eqs. �15�–�17� and other information in Ref. 11.

Repeating this section’s analysis but starting with theassumption that 1 /��T is linear in z instead of the assump-tion of Eq. �21� that � is linear in z leads tediously to

��T�2

TCTH ln�TH/TC��30�

instead of �� /�avg in Eq. �29�. The difference between�T /Tavg and Eq. �30� is only ��T�3 /2Tavg

3 to lowest order in�T /Tavg. The accuracy of the measurements described belowdoes not justify the extra complexity of Eq. �30�, so we retainthe simpler �T /Tavg and �� /�avg dependences in Eqs. �14�and �29�.

The high-amplitude stability of pulse tubes againstgravity-driven convection was characterized by Wang and

8

Gifford in terms of the inverse of the dimensionless group

2278 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

ua2

gD�T/Tavg=

�2a2

gD

Tavg

�T, �31�

which is similar to Eqs. �29� and �14�, but with two impor-tant differences. First, the choice of Ref. 8 keeps g and �Ttogether in the denominator, while our derivation of Eq. �29�shows that the nonlinear nature of the stabilizing effect ofvibrations puts ����2 in the last term in Eq. �27� and, hence,puts �� in the numerator of Eq. �29�, leaving g in the de-nominator: In contrast to the dependence shown in Eq. �31�,higher temperature differences actually allow suppression ofconvection at lower frequencies and amplitudes, even whilea larger acceleration of gravity would require higher ampli-tudes. Second, Ref. 8 arbitrarily chose D as the characteristiclength in the dimensionless group, while our derivationshows that the characteristic length might best be consideredto be � dependent, and that L is more important than D whenD�L, except very close to �=90°.

III. EXPERIMENTS

To investigate these phenomena under a broad range ofexperimental conditions, an apparatus with interchangeabletubes much simpler than complete pulse-tube refrigeratorswas built. Working only at and above ambient temperatureallowed the use of easily measured electric-resistance heat,without refrigeration, and adoption of nearly standing-wavephasing for the measurements eliminated need for a pulse-tube refrigerator’s orifice and compliance tank, simplified theapparatus, reduced surface areas that could contribute toroom heat leaks, reduced the heat demands on the heat ex-changers, and led to rapid thermal-equilibration times. Fivetubes, shown to scale in Fig. 3 and described in Table I, wereused for these measurements.

Each of the five pulse tubes �or thermal-buffer tubes�was a right-circular cylindrical space bounded around itssides by a 0.8-mm-thick stainless-steel wall and on its endsby diffusion-bonded stainless-steel screens acting as flowstraighteners. Each flow straightener comprised 27 layers ofnominally 16.5 wires/cm, 0.14-mm-diameter-wire square-weave screen, with alternate layers turned 45°. They werecut to a diameter that was 1.6 mm greater than that of eachpulse-tube’s inside diameter, by wire electric-discharge ma-chining after diffusion bonding, so steps on the ends of thepulse tube could hold them firmly in place and define thepulse-tube length L accurately. Beyond these flow straight-eners were drilled copper disks that served as heat exchang-ers, maintaining nearly isothermal planes across the ends ofthe flow straighteners by conducting heat to or from theirsurroundings. The heat-exchanger holes were 1.32 mm indiameter, and the hole patterns were designed for spatiallyuniform coverage over the pulse-tube area.

On the hot end, a bounce space the same diameter as thepulse tube allowed significant oscillating flow through thehot heat exchanger, and a 1.5-mm-diameter sheathed type-Kthermocouple in that space, a few millimeters from the hotheat exchanger, measured TH. The thermocouple was bent, asshown in the figure, so almost 1 cm of its tip lay close to the

heat exchanger �except for the thinnest tube, in which the

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum

Page 7: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

1/4 the scale of �a�.

J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009 G. W.

bent portion was necessarily shorter�. A commercial “band”electric-resistance heater provided heat, that heat beingspread around the hot heat-exchanger region by a copperbushing. Ceramic-fiber thermal insulation covered all the hotparts including the pulse tube itself. The cold end wasmounted in a water-cooled aluminum base, whose tempera-ture TC varied no more than 1 °C during the course of anysingle data set, and did not differ from 20 °C by more than afew degrees from week to week.

A passage through the aluminum base, a few centimeterslong, led from the cold heat exchanger to the top of a 10-cm-diameter piston, which was driven by a linear motor14 towhose housing the aluminum base was bolted. The motorwas best operated very near the resonance frequency definedby its large moving mass and the gas-pressure spring con-stant experienced by the piston. This resonance frequencywas easily varied by adjusting the mean pressure, and couldbe changed for a desired mean pressure by inserting volume-adding spacer rings between the motor housing and the alu-minum base. The motor housing was mounted on a modifiedrotary stand, originally intended for rebuilding automobileengines. The rotary part of the stand had a hole-and-pinmechanism for reproducibly setting the tilt of the entire ap-paratus in 10.0° increments from 0.0° to 180.0°. A bubblelevel was used to align the pulse tube with gravity to �0.1°with the apparatus set at 180.0°. The pressure amplitude pa

applied at the bottom of the pulse-tube assembly was mea-sured with a lock-in amplifier connected to a piezoresistivetransducer15 in the aluminum base.

In the D /L=0.52 tube, a second thermocouple was in-stalled, in the copper bushing under the electric-resistanceheater. Near TH=250 °C, the bushing thermocouple wasnever more than 10 °C hotter than the internal thermo-couple, this temperature difference being largest when theconvective heat transport was the largest.

xperiments, and the heat Qgascon carried by simplemental circumstances. See also Fig. 3�c� for scale

9 0.521 0.750 1.57

ns3.64 1.74 3.646.99 2.32 2.320.64 0.64 0.640.82 0.82 0.820.56 0.56 0.56

331 91 3311 0.436 0.521 0.436

0.66 0.66 0.666.74 9.02 9.02

8 0.65 0.45 1.96619

FIG. 3. �a� Cross-sectional scale drawing of the D /L=0.25 tube, shown withthe cold end down �i.e., �=0�. The dimensions Lj are given in Table I. Thepressure-vessel boundary, shown in heavy black, was a long, machined tubewith a cap welded into one end. The cap pressed on a thin-walled sleeve inthe bounce space �not distinct in the figure�, whose inside diameter was thesame as that of the pulse tube, thereby trapping the hot heat exchanger andhot flow straightener against a machined step at the top of the pulse tube.The cap included a welded-in compression fitting through which the hotthermocouple passed. Clamps holding the tube to the base are not shown.�b� A perpendicular cross section though one of the drilled copper heatexchangers, at twice the scale of �a�. �c� The proportions of all five tubes, at

TABLE I. Dimensions for the five tubes used in the econduction in the gas in each tube under typical experidrawings.

D /L= 0.126 0.24

DimensioD �cm� 0.88 1.74L �cm� 6.98 6.99Lfs �cm� 0.64 0.64� fs 0.82 0.82Lhx �cm� 0.56 0.56No. of hx holes 19 91�hx 0.427 0.52Holes’ Rhx �mm� 0.66 0.66Lbounce �cm� 6.66 6.74

Qgascon �WHe, TH=425 °C 0.30He, TH=250 °C 0.038 0.14He, TH=150 °C 0.070.9He–0.1Ar, TH=250 °C 0.12Ar, TH=250 °C 0.01

Swift and S. N. Backhaus: The pulse tube and the pendulum 2279

Page 8: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

Obtaining one set of data typically took half of a day. Agas, its mean pressure pm, a frequency, and a tip angle � werechosen, and were kept fixed for each data set. An initialmotor drive voltage was chosen, and heat was applied to theelectric-resistance heater to maintain the hot temperature at aselected TH. To assess that process, temperature was dis-played as a function of time with a chart recorder. The heatervoltage was adjusted manually until a steady settingachieved both a low rate of change in temperature �less than0.1 °C in a few minutes� and the desired TH. The steady-state heater voltage V and pressure amplitude pa were then

recorded, and the heater power Q was obtained by squaringthe voltage and dividing the result by the heater’s resistance.The motor drive voltage was changed to a new value, and theheat adjustment and data recording were repeated, typicallyat rates of two to four data points per hour.

Figure 4�a� shows six such data sets, all in the D /L=0.25 tube with 3.1-MPa helium gas driven at 100 Hz, andwith TC=20 °C and TH=250 °C.

At �=0°, the cold end of the pulse tube was straightdown so the gas was convectively stable. The measurementsshow that 14 W of heat were needed to keep TH=250 °C inthis tube, with amplitude-dependent variation being only afraction of 1 W. Calculations show that the helium in thepulse tube conducted 0.15 W and the stainless-steel pulse-tube wall conducted 2.5 W, so most of the required heat wasapparently heat leak through the fiber insulation to the room.Calculations16 that include boundary-layer heat shuttle alongthe pulse tube and acoustic-power dissipation in the hot heatexchanger and flow straightener show that the required heatshould drop quadratically by 0.8 W as the oscillation ampli-tude rises from pa / pm=0.01 to pa / pm=0.05, in rough agree-ment with the �=0° measurements.

Compared with �=0°, only a little more heat was con-vected at �=60° at zero or low oscillation amplitude. Thistube was slender enough that the highest edge of its cold endwas still 1.0 cm below the lowest edge of its hot end at �=60°, so the gas in the tube can still be regarded as convec-tively stable at this tip angle.

At �=90°, 120°, 150°, and 180° over 4 W of heat wasconvected through the tube when no oscillations werepresent, representing Nusselt numbers ranging from 30 at90° to 50 at 150°. Such convection is large enough to reducethe cooling power of a pulse-tube refrigerator significantly.The Rayleigh number based on L is about 27�106, and suchNusselt numbers are plausible at this Rayleigh number: Eq.�4.89� in Ref. 17 yields a Nusselt number of 18 under theseconditions, for �=180°. �However, our enclosure has porousends, which could tend to increase the Nusselt number.�From the convective heat flow, �, cp, and �T, we estimatethat the convective velocity was of the order of 1 cm/s,roughly 100 times less than the typical oscillating velocity.The Reynolds number of the convective motion here is of theorder of 20, so the convection should be laminar. This sug-gests that numerical calculations based on Ref. 11 may yieldreliable results in this range of parameter space. However, inthe tubes with D /L�0.5 we did sometimes see time-dependent convection, evidenced by time dependence in the

hot temperature, whose variations were as high as a few

2280 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

tenths of a degree over time scales of about 10 s. The timedependence started near the convection-suppression transi-tion and rose with amplitude, and was greatest for the tubewith the highest D /L. Numerical calculations in the time-dependent regime might be more challenging.

Figure 4�a� shows an initial rise in convective heat trans-fer with amplitude for ��120°. Possible explanations forthis phenomenon include a weakening of the zero-velocityboundary condition at the ends of the convective cell asthose ends find themselves, on average, farther from the flowstraighteners at higher amplitude, and a strengthening of

FIG. 4. Heat Q required to maintain a steady hot temperature under a widevariety of conditions with 3.1-MPa gas in the D /L=0.25 tube. The pointsare measurements, and the lines are only guides to the eyes. �a� Heat re-quired to maintain TH=250 °C, with helium at 100 Hz, for six tip angles �.The horizontal axis is the pressure amplitude at the base, divided by meanpressure. �b� Heat required to maintain TH=250 °C, with helium at �=120°, for three frequencies. �c� Heat required to maintain TH=250 °C, at100 Hz, at �=120°, for four different gases. The mixture was 90% heliumand 10% argon. The horizontal axis in �b� and �c� is the square root of therelevant part of Eq. �14�, because this keeps the points almost equallyspaced, making the variations near the transition easier to see.

thermal contact near the ends of the convective cell as jets

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum

Page 9: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

whose diameters are of the order of the flow-straighteners’hydraulic radius squirt gas at the heat-exchangers’ tempera-tures into the ends of the convective cell.

Figure 4�a� also shows that the convection was effec-tively stopped when the oscillations had a high enough am-plitude, as expected from Secs. I and II.

The closely spaced points near the 120° transition andthe essentially overlapping points throughout the 180° dataindicate attempts to observe hysteresis, by taking some of thedata while systematically increasing pa and other data whilesystematically decreasing pa. No hysteresis was observed inthese data sets or in any others. �A near exception is de-scribed in the caption for Fig. 6�b�.�

To plot such data as a function of �a or of Nptc, weconverted from the measured pa / pm to the vibration ampli-tude a in the middle of the pulse tube by using

a =pa

�pmL

2+ �� fsLfs + 1 + �� − 1�

��

Rhx��hxLhx + Lbounce ,

�32�

where half of the pulse-tube length, L /2, the hot-flow-straightener length Lfs, the hot-heat-exchanger length Lhx,and the bounce-space length Lbounce add up to the total dis-tance between the middle of the pulse tube and the closedend of the experiment. This expression is based on the as-sumptions that pa is independent of x from the middle of thepulse tube to the end of the bounce space, and that thermal-hysteresis effects in the bounce space and pulse tube arenegligible. Thus, if the total distance were unobstructed andof uniform cross-sectional area, then Eq. �32� would be sim-ply a= �pa /�pm��L /2+Lfs+Lhx+Lbounce�, describing simpleadiabatic compressions and expansions everywhere. Theprefactor pa /�pm is used in Eq. �32� because most of thatlength, L /2+Lbounce, does experience nearly adiabatic com-pressions and expansions. The prefactors of the small Lfs andLhx terms in Eq. �32� account for the volumetric porosities � j

of those components, the isothermal nature of the oscillationsin the flow straightener, and the thermal hysteresis in thecircular channels through the heat exchanger, in which thechannel radius is Rhx and the thermal penetration depth is ��.Numerical estimates16 that include inertial and resistive pres-sure drops in the hot heat exchanger and flow straightenerand the consequences of thermal hysteresis elsewhere sug-gest that these assumptions introduce errors of no more than2% to the determination of a.

Figure 4�b� shows convection-suppression data from theD /L=0.25 tube at three different frequencies, all with 3.1-MPa helium at �=120° and TH=250 °C. Although the fre-quency ranges over a factor of 2, plotting these data sets asfunctions of �a aligns them very well, corroborating the �afunctional dependence on Nptc in Eq. �14�, and contradictingany other supposed strong dependences on � or a in thistube, such as a /L �independent of �� or �a2 /L.

�The “100 Hz �orig�” data set in Fig. 4�b� is also shownin Fig. 4�a�. After taking that data set and the “60 Hz �orig�”set, the original D /L=0.25 tube was disassembled to useparts elsewhere, and later was “rebuilt” to obtain more data.

The difference between the “original” and “rebuilt” 100-Hz

J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009 G. W.

sets is presumably due to slight hardware irreproducibility,including slightly different hot-thermocouple positions. Forfuture work, we recommend a reproducible attachment ofboth thermocouples directly to their copper heat exchangers.�

Figure 4�c� shows convection-suppression data from theD /L=0.25 tube for four different gases at 3.1 MPa and 100Hz, with TH=250 °C and �=120°. The horizontal alignmentof all of these data sets confirms the lack of explicit gas-property dependence of Nptc. The alignment of the heliumand argon data, despite the tenfold difference in atomic massand mass density, confirms that Nptc should be independentof molecular mass. The alignment of the helium-argon datawith the pure-monatomic-gas data, despite the mixture’s32% lower Prandtl number, confirms that Nptc is independentof Prandtl number and, by inference, independent of the gastransport properties. The alignment of the �=7 /5 nitrogendata with the �=5 /3 monatomic-gas data confirms that Nptc

is independent of the specific-heat ratio, except through theconversion from pa / pm to a given in Eq. �32�.

To investigate the �T dependence of the convection-suppression transition, we used the D /L=0.25 tube with 3.1-MPa helium and �=120° at 100 Hz, at three different hottemperatures. To bring the data into approximate vertical

alignment, we divided Q by �T, and then subtracted 0.08,0.06, and 0.05 W / °C from the 425, 250, and 150 °C data,respectively, to account for the temperature-dependent heatleaks. With the three data sets plotted against �a in Fig. 5�a�,it is apparent that it was easier to suppress the convection athigher �T. Figures 5�b� and 5�c� show these three data setsplotted against �a�4 �T /Tavg and �a��T /Tavg. The data alignbest using the fourth root, which is why we choose =1 /2 inEq. �14�, despite the fact that the derivation of Sec. II yields=1.

To study the L and D dependence of the convection-suppression transition, we used data from all five pulse tubes,which had five different aspect ratios. Measurements withidentical gas, temperatures, and frequency are shown in Figs.4�a� and 6. Like the D /L=0.25 tube, which yielded the datashown in Figs. 4 and 5, the D /L=0.52 tube displayed sharpconvection-suppression transitions at �=120° and 150°, anda �=0 heat requirement that was almost independent of am-plitude, as shown in Fig. 6�b�. In the D /L=0.75 tube, thetransitions were still sharp, but the �=0 heat requirementrose dramatically, and quadratically, with amplitude, asshown in Fig. 6�c�. The D /L=1.57 tube showed a similarrising baseline heat requirement, but a very ill-defined andincomplete transition to reduced convection, as shown inFig. 6�d�. Our motor did not let us learn whether higheramplitudes would bring a second, more complete reductionin convection in this tube. Unlike the other four tubes, theD /L=0.126 tube did not have sharp transitions at any tipangle, as shown in Fig. 6�a�, and the small heats involvedwere difficult to measure accurately.

We do not understand some of these qualitative differ-ences between the data sets in the different tubes. The qua-dratically rising �=0 heats for the two shortest tubes suggeststreaming, but the calculated Rayleigh streaming velocity18

just outside the boundary layer at mid-tube is very nearly the

same, 1.3 cm/s at pa / pm=0.025, for all five tubes, and esti-

Swift and S. N. Backhaus: The pulse tube and the pendulum 2281

Page 10: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

mates of the heat that such streaming can transport along thetubes19 range only from 0.1 to 0.5 W at that pa / pm, too smallto explain the measurements. Seeking another reason that theshort tubes differ from the long tubes, one can consider thestroke divided by the tube length, 2a /L, which should besmaller than 1 to prevent gas from shuttling heat all the wayfrom the hot flow straightener to the cold flow straightenerevery cycle of the oscillation. But 2a /L is only 0.12 at the�=120° transition in the D /L=0.75 tube, where the risingbaseline is perhaps even visible as low as 2a /L�0.07, soshuttle heat should not be responsible for the rising baseline.Furthermore, the D /L=0.25 tube’s 100-Hz, �=0 data reachas high as 2a /L=0.09, and that tube’s 45-Hz data extend to2a /L=0.14, with no suggestion of rising baselines in Figs.

FIG. 5. Normalized heat required to maintain a steady hot temperature inthe �rebuilt� D /L=0.25 tube, with 3.1-MPa helium and �=120°, at threedifferent hot temperatures TH with which the points are labeled. �N.B.: �T is�20 °C smaller than TH.� The points are measurements, and the lines areonly guides to the eyes. The experimental temperature dependence is closerto the fourth root used in �b� than to either the square root used in �c� or notemperature dependence used in �a�.

4�a� or 4�b�.

2282 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

Despite this mystery, the data from the four tubes withthe smallest D /L can be used to explore whether D or L isthe most important geometrical variable l in Eq. �20� govern-ing the convection-suppression transition, and to test the�D sin �−L cos � geometry dependence given in Eq. �29�for small D /L. From each ��0 data set in Figs. 4�a� and 5,

FIG. 6. Heat required to maintain TH=250 °C in 3.1-MPa helium at 100Hz, in four different tubes. Figure 4�a� shows similar data for a fifth tube.The points are measurements, and the lines are only guides to the eyes. Thehorizontal axis is pressure amplitude at the base, divided by mean pressure.In �c�, one of the 180° points is not connected with the lines. That data pointwas metastable: It persisted steadily for 5 min before the heat suddenlydropped to the point below it.

we subtracted a quadratic fit to the corresponding �=0 base-

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum

Page 11: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

line data set, and defined the transition from convection tosuppression as the value of pa / pm where each data set passes

halfway between the maximum value of Q− Qbaseline andzero. �This definition of the transition is essentially identical

to the location of steepest decrease in Q− Qbaseline, except forthe D /L=1.57 tube.� We converted the transition value ofpa / pm to a corresponding transition value of a using Eq.�32�.

Figure 7 displays the results as a function of D /L forthree different choices of the characteristic length l thatmight be used in the dimensionless group in Eq. �20�. First,Fig. 7�a� shows the results when plotted with l=D, thechoice made in Ref. 8. For this choice of l, the transitiondisplays complicated dependence on � and D /L, suggestingthat simply using l=D in Nptc does not provide a universaldescription of the transition. In fact, for �=180°, the transi-tion varies almost as 1 / �D /L�, as shown by the dashed curve,suggesting that dividing by D is a very poor choice for thisparticular �. Next, in Fig. 7�b�, the same data are plottedusing l=L. Here, the ��90° data collapse reasonably wellalong a single curve with little D /L dependence, but the �=90° data deviate significantly from the others; comparisonto the dashed line shows that the �=90° transition variesalmost as D /L for small D /L, suggesting that dividing by Lis a poor choice for this particular �. Finally, Fig. 7�c� showsthe same data plotted using l=�D sin �−L cos �, with �=1.5. This choice brings the data sets for all tip angles closeto a common curve, consistent with Eq. �29� in some ways.Trying �=1.0, 2.5, or more-extreme values ruins the clus-tering of the data in Fig. 7�c�, while using �=2.0 looks onlya little worse than �=1.5. Using �=1.5 sets �slow=2L /3��=0.14L, which seems reasonable, being about three timeslarger than the �slow shown in Fig. 2.

IV. FURTHER DISCUSSION

The vibrational stabilization of an inverted pendulum isa useful guide to intuition about how acoustic oscillationssuppress natural convection in an inverted pulse tube, andthe dimensionless pulse-tube convection number Nptc definedin Eq. �14� may provide a good quantitative framework foranalysis, at least for small aspect ratios. Experiments confirmthat �a captures the relevant dependences on frequency anddisplacement, and that gas properties such as � and Prandtlnumber are not important. However, the picture is incom-plete, at best.

For example, the observed =1 /2 temperature depen-dence in Fig. 5 differs significantly from the =1 predictionof Eq. �29�. This remains a mystery. In the same figure, di-

viding Q by �T brought the data into good vertical align-ment, implying that the Nusselt number is independent of�T, while Eq. �4.89� in Ref. 17 predicts that the Nusseltnumber should be proportional to ��T�1/3.

Furthermore, we are not sure how to interpret the D /Ldependence that remains in Figs. 7�c� and 7�d�. One possi-bility is that the transition occurs at Nptc�1 for a substantialrange of D /L, including 0.25�D /L�0.52, as predicted forlow D /L by Eq. �14� and suggested by the dashed line in Fig.

7�c�. The data at D /L=0.126 might fall below this value

J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009 G. W.

because of physics not included in Sec. II: For example, inthe D /L=0.126 tube, transverse thermal relaxation is fasterthan in any other tubes, and is comparable to the �l /gconvective-motion time, both because the helium-column di-ameter is smaller and because of the relatively greater con-tribution of circumferential conduction by the stainless-steel

FIG. 7. Transitional values of Nptc based on data from Figs. 4�a� and 6,testing different choices for l in Eq. �20�. �a� Choosing l=D yields transi-tional values of Nptc that are not independent of aspect ratio at small aspectratio when ��90°. The dashed curve shows Nptc�1 / �D /L�, for comparisonwith the �=180° data. �b� Choosing l=L yields transitional values of Nptc

that are more nearly independent of aspect ratio at small aspect ratio when��90°, but leaves the �=90° data varying almost linearly with aspect ratio,as suggested by the dashed line. �c� Choosing l=�D sin �−L cos �, with�=1.5, yields transitional values of Nptc that are more nearly independent ofboth � and D /L. �d� For completeness, this shows the same data and verticalaxis as in �c�, but with the data from the highest-aspect-ratio tube included.

tube wall. A second interpretation would simply discount the

Swift and S. N. Backhaus: The pulse tube and the pendulum 2283

Page 12: The pulse tube and the pendulum · 2018-11-28 · pendulum is vibrated at high enough frequency and high enough amplitude, the pendulum tends to align with the vi-bration axis, and

D /L=0.126 data because the convection there was weak,differed qualitatively from the other data sets �having no ini-

tial rise of Q with pa�, and was hard to measure well �e.g.,day-to-day and hour-to-hour variations in room temperaturewould have had a greater effect on this data set than on theothers�. A third interpretation would be that the analysis ofSec. II is wrong and the data show that Nptc�c1+c2D /Ldescribes the transition for all D /L, with the data at D /L=0.52 to be discounted for some unknown reason.

Resolving these and the other interesting, unansweredquestions raised here may require additional experiments,numerical modeling, or both. One important question iswhether the suppression of convection depends on the oscil-lating pressure at all. The time phase difference between theoscillating pressure and oscillating velocity in these experi-ments was 90°, while practical pulse tubes, transmittingacoustic power, operate closer to a time phase of 0° or 180°,in some cases with the time phase tuned to reduce Rayleighstreaming.18 Whether this time phase affects the convectionsuppression, either directly or indirectly through Rayleighstreaming, has not been investigated here. And the magni-tude of the oscillating pressure, neglected here in the discus-sion between Eqs. �20� and �21�, might have a significanteffect via the gas compressibility, because it makes the os-cillating velocity at the ends of the pulse tube different fromthat in the center.

This situation is most unclear for D /L�1, where Fig.6�d� shows that the suppression of convection by vibration isvery incomplete, or, at best, only partially completed at theamplitudes accessible in this experiment. At such aspect ra-tios, and with a sometimes a significant fraction of L, ensur-ing that imperfect flow straightening at the ends of the tubedoes not affect the measurements may be particularly chal-lenging.

Other well-known rigid-pendulum phenomena, such asparametric resonance and synchronized unidirectionalrotation,1 may also have analogs in pulse tubes, at lowerfrequencies than those studied here.

ACKNOWLEDGMENTS

This work was supported by license income from ther-moacoustic patents at Los Alamos National Laboratory. Wethank Robert Keolian and Robert Ecke for many insightsabout the effects of vibrations on pendula and fluids, and forintroductions to the relevant literature.

2284 J. Acoust. Soc. Am., Vol. 126, No. 5, November 2009

1A recent summary and introduction to some of the relevant literature aregiven by E. I. Butikov, “On the dynamic stabilization of an inverted pen-dulum,” Am. J. Phys. 69, 755–768 �2001�.

2L. Blitzer, “Inverted pendulum,” Am. J. Phys. 33, 1076–1078 �1965�.3L. D. Landau and E. M. Lifshitz, Mechanics �Pergamon, New York, 1960�,Sec. 30.

4D. Morin, Introduction to Classical Mechanics �Cambridge UniversityPress, Cambridge, 2008�, Solution 6.5.

5R. Radebaugh, “Development of the pulse tube refrigerator as an efficientand reliable cryocooler,” Proceedings of the Institute of Refrigeration,London �2000�, pp. 11–29; http://cryogenics.nist.gov/Papers/Institute_of_Refrig.pdf �Last viewed 9/30/2009�.

6G. Thummes, M. Schreiber, R. Landgraf, and C. Heiden, “Convective heatlosses in pulse tube coolers: Effect of pulse tube inclination,” in Cryocool-ers 9, edited by R. G. Ross �Plenum, New York, 1997�, pp. 393–402.

7G. Thummes and L. Yang, “Development of Stirling-type pulse tube cool-ers driven by commercial linear compressors,” in Infrared Technology andApplications XXVIII, edited by B. F. Andresen, G. F. Fulop, and M.Strojnik �Society of Photo-Optical Instrumentation Engineering, Seattle,WA, 2003�, pp. 1–14.

8C. Wang and P. E. Gifford, “A single-stage pulse tube cryocooler forhorizontally cooling HTS MRI probe,” in Advances in Cryogenic Engi-neering: Transactions of the Cryogenic Engineering Conference—CEC,edited by J. Waynert, J. Barclay, S. Breon, E. Daly, J. Demko, M. DiPirro,J. R. Hull, P. Kelley, P. Kittel, A. Klebaner, J. Lock, J. Maddocks, J.Pfotenhauer, C. Rey, and Q.-S. Shu �American Institute of Physics, NewYork, 2004�, Vol. 49, pp. 1805–1811.

9S. Backhaus and G. W. Swift, “A thermoacoustic-Stirling heat engine:Detailed study,” J. Acoust. Soc. Am. 107, 3148–3166 �2000�.

10D. L. Gardner and G. W. Swift, “A cascade thermoacoustic engine,” J.Acoust. Soc. Am. 114, 1905–1919 �2003�.

11G. Z. Gershuni and D. V. Lyubimov, Thermal Vibration Convection�Wiley, New York, 1998�.

12G. W. Swift, Thermoacoustics: A Unifying Perspective for Some Enginesand Refrigerators �Acoustical Society of America, Sewickley, PA, 2002�.

13G. W. Swift, Thermoacoustics: A Unifying Perspective for Some Enginesand Refrigerators �Ref. 12�, Eq. �7.53�.

14Model C2, Clever Fellows Innovation Consortium, Inc., Troy, NY, http://qdrive.com �Last viewed 9/30/2009�.

15Model 8510B-500M37, Endevco, San Juan Capistrano, CA, www.endev-co.com �Last viewed 9/30/2009�.

16B. Ward, J. Clark, and G. Swift, “Design environment for low-amplitudethermoacoustic energy conversion �DeltaEC�,” software and user’s guideavailable from the Los Alamos thermoacoustics web site http://www.lanl.gov/thermoacoustics/ �Last viewed 9/30/2009�.

17W. M. Rohsenow, J. P. Hartnett, and Y. I. Cho, Handbook of Heat Transfer�McGraw-Hill, New York, 1998�.

18J. R. Olson and G. W. Swift, “Acoustic streaming in pulse tube refrigera-tors: Tapered pulse tubes,” Cryogenics 37, 769–776 �1997�.

19K. I. Matveev, S. Backhaus, and G. W. Swift, “The effect of gravity onheat transfer by Rayleigh streaming in pulse tubes and thermal buffertubes,” in Proceedings of the IMECE 2004: International Mechanical En-gineering Congress and Expo, Fairfield, NJ �American Society of Me-chanical Engineers �ASME�, New York, 2004�, pp. 125–139, Paper No.IMECE 2004-59076.

G. W. Swift and S. N. Backhaus: The pulse tube and the pendulum