• No se han encontrado resultados

Determination of Maximum Mechanical Energy Efficiency in Energy Galloping Systems

N/A
N/A
Protected

Academic year: 2020

Share "Determination of Maximum Mechanical Energy Efficiency in Energy Galloping Systems"

Copied!
9
0
0

Texto completo

(1)

1

Determination of Maximum Mechanical Energy Ef

ciency in

2

Energy Galloping Systems

1

3

J. Meseguer

1

; A. Sanz-Andrés

2

; and G. Alonso

3

4 Abstract:Transverse galloping is a type of aeroelastic instability characterized by large amplitude, low frequency, normal to wind oscillations 5 that appear in some elastic two-dimensional bluff bodies when subjected to afluidflow, provided that theflow velocity exceeds a threshold 6 critical value. Such an oscillatory motion is explained because of the energy transfer from theflow to the two-dimensional bluff body. The 7 amount of energy that can be extracted depends on the cross section of the galloping prism. Assuming that the Glauert-Den Hartog quasi-8 static criterion for galloping instability is satisfied in afirst approximation, the suitability of a given cross section for energy harvesting is eval-9 uated by analyzing the lateral aerodynamic force coefficient,fitting a function with a power series in tana(abeing the angle of attack) to 10 available experimental data. In this paper, a fairly large number of simple prisms (triangle, ellipse, biconvex, and rhombus cross sections, as well 11 as D-shaped bodies) is analyzed for suitability as energy harvesters. The influence of thefitting process in the energy harvesting efficiency 12 evaluation is also demonstrated. The analysis shows that the more promising bodies are those with isosceles or approximate isosceles cross 13Q : Asections.DOI:10.1061/(ASCE)EM.1943-7889.0000817.©2014 American Society of Civil Engineers.

14 Author keywords:Galloping; Energy harvesting; Wind tunnel.

15 Introduction

16 The potential use of different aeroelastic phenomena to extract en-17 ergy from an incidentow, and through the oscillations of the ex-18 cited structure, convert those oscillations into piezoelectric or 19 electromagnetic energy, is receiving the attention of more and more 20 researchers (St. Clair et al. 2010;Peng and Chen 2012). As a con-21 sequence, numerous studies have been published recently on the use 22 offluid-bodies instabilities interactions (the number of papers has 23 remarkably grown in the last 2 years). The focus of these papers has 24 been either on the particularities of oscillating systems (transverse 25 galloping, wake galloping,flutter, and vortex shedding) or on the use 26 of these oscillating systems for energy harvesting.

27 Besides the pioneer work of Glauert (1919), Den Hartog (1965), 28 Novak (1969), Novak and Tanaka (1974), and Luo et al. (1988), 29 among others, some relevant publications have appeared more re-30 cently concerning transverse galloping instability (Naudascher and 3119 Rockwell 1994;Luo et al. 1998;Barrero-Gil et al. 2010; Païdoussis 32 et al. 2010,Sewatkar et al. 2012,Alonso and Meseguer 2014;Ibarra 33 et al. 2014), whereas the use of transverse galloping for energy 34 harvesting purposes has been analyzed by Sirohi and Mahadik (2011, 35 2012), Abdelkefiet al. (2012a,2013a), Abdelkefiet al. (2013c,2014), 36 and Zhao et al. (2012).

37 However, transverse galloping is not the onlyflow-induced os-38 cillating phenomenon in energy harvesting devices, and some

39 efforts have been devoted to using other instability sources, such as

40 wake galloping (Jung et al. 2009;Akaydin et al. 2010;Jung and

41 Lee 2011;Abdelkefiet al. 2013b,c), vortex-induced vibrations

42 (Abdelkefiet al. 2012b;Grouthier et al. 2012;Hobbs and Hu 2012;

43 Peng and Chen 2012;Mackowski and Williamson 2013;Mehmood

44 et al. 2013), andflutter of streamlined bodies (Peng and Zhu 2009;

45 Tang et al. 2009;Zhu et al. 2009;Erturk et al. 2010;Bryant et al.

46 2011;Doaré and Michelin 2011;Zhu 2011;Abdelkefiet al. 2012c,d;

47 Abdelkefiand Nuhait 2013;Bae and Inman 2014).

48 Most of these studies focused on the analysis of the suitability of

49 some specific mechanisms to efficiently harvest the energy through

50 different energy converters (piezoelectric, electromagnetic, etc.). In

51 contrast, some attempts have been made to develop methodologies

52 to evaluate the energy harvesting efficiency of two-dimensional

53 bluff bodies from a purely aeroelastic point of view.

54 In this sense, Barrero-Gil et al. (2010) developed a simple

ap-55 proach to evaluate the efficiency of the transverse galloping of bluff

56 bodies as an energy harvesting mechanism. To do this, following an

57 already published methodology (Ng et al. 2005), afirst

approxi-58 mation of the transverse galloping phenomenon can be explained

59 by a one-degree of freedom (the transverse displacement,z) linear

60 differential equation. The forcing aerodynamic coefficientczis

61 expanded in a power series of the transverse velocity, dz=dt,

62 retaining for simplicity only, the twofirst terms in theczseries

63 expansion, that is,cz5a1ðdz=dtÞ=U1a3½ðdz=dtÞ=U3. Barrero-Gil

64 et al. (2010) demonstrates that, within this approximation, the

65 maximum harvesting efficiency ishmax5 2a2

1=ð6a3Þ. The criterion

66 developed is applied to assess the energy harvesting efficiency of

67 several two-dimensional bodies (square, isosceles triangle, and

68 D-section) whose aerodynamic coefficients are taken from the

69 existing literature. The conclusion drawn is that D-shaped,

two-70 dimensional bodies are the most appropriate ones for energy

har-71 vesting devices based on transverse galloping, because they provide

72 the highesthmaxvalue among the considered bodies.

73 Further experiments conducted at IDR/UPM on families of

74 bodies with different cross sections provided a better insight on the

75 transverse galloping phenomenon, and allowed for improvement of

76 the approach initiated by Barrero-Gil et al. (2010) to predict the

1Professor, IDR/UPM, E.T.S.I. Aeronáuticos, Univ. Politécnica de

Madrid, E-28040 Madrid, Spain. E-mail: j.meseguer@upm.es

2

2Professor, IDR/UPM, E.T.S.I. Aeronáuticos, Univ. Politécnica de

Madrid, E-28040 Madrid, Spain. E-mail: angel.sanz.andres@upm.es 3Professor, IDR/UPM, E.T.S.I. Aeronáuticos, Univ. Politécnica de

Madrid, E-28040 Madrid, Spain (corresponding author). E-mail: gustavo .alonso@upm.es

Note. This manuscript was submitted on May 20, 2013; approved on May 2, 2014; published online onnnnnnn. Discussion period open until

nnnnnnnnnnnnnnnnnn; separate discussions must be submitted for in-dividual papers. This paper is part of theJournal of Engineering

(2)

77 energy harvesting efficiency of these structures by also providing 78 a larger set of experimental aerodynamic data to validate the meth-79 odology. In particular, the transverse galloping stability of triangles 80 (Alonso et al. 2005,2007;Alonso and Meseguer 2006), ellipses 81 (Alonso et al. 2010), biconvex, and rhombi (Alonso et al. 2009) were 82 thoroughly determined by wind tunnel testing of cylinders with 83 respective cross-sectional geometry and variable relative thickness. 84 By properly using the new and extended set of experimental 85 aerodynamic data, it is possible to further develop and improve the 86 methodology presented in Barrero-Gil et al. (2010) to evaluate the 87 transverse galloping energy efficiency of bluff bodies.

88 A brief summary of the theoretical foundation already published 89 in Barrero-Gil et al. (2010) is provided in theMathematical Model 90 section, and a new and improved way to estimateh

maxis presented.

91 The section Analysis of the Energy Harvesting Properties of 92 Several Bluff Bodiesis devoted to the application of theh

max

93 criterion to a large amount of new cross-sectional shapes. Con-94 clusions are outlined in theConclusionssection.

95 Mathematical Model

96 Theoretical foundations of galloping are well established and un-97 derstood (Den Hartog 1965; Novak 1969, 1972). The simplest 98 model to analyze the translational galloping behavior of a bluff body 99 consists of a damped spring-mounted cylindrical body under the 100 action of an incoming smooth flow of velocityU, parallel to the 101 x-axis (Fig.1). Other relevant magnitudes are the cylinder mass per 102 unit length,m, the mechanical damping ratio,z, the natural circular 103 frequency of oscillations,vn, thefluid density,r, and a characteristic 104 transversal length of the galloping two-dimensional body (as the 105 normal to wind length of its cross section,bm). Then, assuming the 106 body is slender enough to assure two-dimensional ow, the dy-107 namics of the system is driven by

m

d2z dt2þ2zvn

dz dtþv

2

nz

¼ fzðaÞ ¼1 2rU

2b

mczðaÞ (1)

108 wherez5vertical displacement;t5time;f

z5instantaneous two-109 dimensional lateral force (parallel to thez-axis);cz5instantaneous 110 two-dimensional lateral force coefcient; vn 5 undamped reso-111 nance frequency; andz5damping coefficient.

112 In a body reference frame, across-wind oscillation of the structure

113 periodically changes the angle of attack of the incident wind. Such

114 a variation in the angle of attack produces variation in the

aero-115 dynamic forces acting on the structure (Fig.1), which, in turn, affects

116 the dynamic response of the structure. Therefore, transverse

117 galloping can be explained as although the incident velocityUis

118 uniform and constant, because of the lateral oscillation of the body,

119 in a body reference system, the total velocity changes both its

mag-120 nitude and direction with time. Consequently, the body angle of attack

121 also changes with time, hence, the aerodynamic forces acting on it.

122 Consider that a structure at rest is oriented at a given angle of

123 referencea0with respect to the incidentflow (Alonso et al. 2005;

124 Alonso and Meseguer 2006). When the body oscillates along

125 the z-axis direction within an uniform flow with velocity U,

126 the relative velocity between the fluid and the body is Ur

127

5U21ðdz=dtÞ21=2

5U11tan2a1=2, where the angle of

at-128 tack caused by the oscillation is tana5ðdz=dtÞ=U; therefore,

129

Ur5U=cosa. Drag,dðaÞ, and lift,lðaÞ, forces are, respectively,

130

dðaÞ5ð1=2ÞrUr2bmcdðaÞ and lðaÞ5ð1=2ÞrU2rbmclðaÞ, where 3 131

cdðaÞ stands for the drag coefficient,clðaÞ is the lift coefficient,

132 ris thefluid density, andbmis the characteristic transversal length

133 of the cylinder, as already stated.

134 The projection of those forces in the z-axis direction is fzðaÞ

135

5edðaÞsinaelðaÞcosa, and then Eq.(1)reads as

m

d2z dt2þ2zvn

dz dtþv

2

nz

¼21

2rU 2

rbm½cdðaÞsinaþclðaÞcosa

¼21

2rU 2b

mcdðaÞtanaþclðaÞ cosa ¼1

2rU 2b

mczðaÞ

(2)

136 To solve Eq.(2), the functionczðaÞcan be expanded in a power

137 series of tana(Barrero-Gil et al. 2010;Ng et al. 2005) byfitting the

138 function czðaÞ, measured, for instance, in wind tunnel tests, to

139 a polynomial in tanaas

czðaÞ ¼P‘

j51

a2j21tan2j21a (3)

140 Only the antisymmetric part ofczðaÞaroundczða0Þcontributes to

141 galloping. Therefore, only the antisymmetric part ofczðaÞis used to

142 obtain the polynomial Eq.(3). As long asais small enough, the linear

143 term coefficienta1 reduces to the function dcl=da1cdappearing

144 in the Glauert-Den Hartog quasi-static criterion for galloping.

145 Considering the movement is quasi-steady, because tana1,

146 the lateral force coefficient becomesczðaÞ≅eðcda1clÞ1Oða2Þ

147

5 2ðcd1dcl=daÞa1Oða2Þ; hence,a15eðcd1dcl=daÞ.

148 The usual dimensionless variables in aeroelastic analysis are now

149 introduced:y5z=bm,t5vnt,mp5m=ðrb2mÞ, andUp5U=ðvnbmÞ

150 in Eq.(2), and substitutingczðaÞby its series expansion in Eq.(3),

151 Eq.(2)yields

d2c dt2þ2z

dc dtþc¼

Up2 2mp

P‘

j51

a2j21 1 Up

dc dt 2j21

(4)

152 4 wherey5dimensionless vertical displacement;t5dimensionless

153 time;mp5ratio of the mean density of the body to the density of the

154 surroundingfluid; andUp5reduced velocity.

155 Once the problem formulation has been established (the

pre-156 ceding differential equation with suitable boundary and initial

157 conditions), an approximation to the problem solution can be

Fig. 1.Two-dimensional body in galloping oscillation:clðaÞ,cdðaÞ, lift and drag coefficients, respectively;czðaÞ, vertical force coefficient; D, damper; dz=dt, vertical velocity caused by galloping motion; S, spring;U, upstream unperturbedflow velocity;a, angle of attack,a0, reference angle

(3)

158 obtained by applying, for instance, the procedure already presented 159 in Barrero-Gil et al. (2010). The method consists of the application of 160 the Krylov-Bogoliubov method to this differential equation to obtain 161 the time variation of the dimensionless galloping amplitude. To

162 solve the problem, it is assumed that the body motion can be

163 expressed as c5ApðtÞcos½t1fðtÞ, whereApðtÞ and fðtÞ are

164 functions slowly varying witht. Then, the dimensionless amplitude

165 becomes

dAp dt ¼2

1 2p ð 2p 0 "

22zdc dtþ

Up2 2mp

P

j51 a2j21 Up2j21

dc dt

2j21#

sinðtþfÞdt

¼zAp2Up2 4mp a1

Ap Upþ

3 4a3

Ap Up

3

þ5 8a5

Ap Up

5

þ35 64a7

Ap Up

7

þ 63 128a9

Ap Up 9 þ/ " # (5)

166 The steady amplitude of the oscillations is given by the real and 167 positive roots of dAp=dt50. Independent of the number of terms 168 retained in Eq.(5), one of the roots is alwaysAp50. The analytical 169 calculation of the other roots of Eq. (5)becomes more and more 170 complicated as the number of terms increases. If onlya1anda3are 171 retained, the efciency, or conversion factor, dened by the ratio of 172 the power extracted by the structure from theflow per unit length 173 and the total power in the flow per unit length, becomes h

max

174 5 2a2

1=ð6a3Þ, as already stated.

175 The power that can be obtained in a cycle from transverse gal-176 loping oscillations,P

g, is obtained from the expression

Pg¼1

T ðT

0 fzdz

dtdt¼

rU2b 2T

ð T

0 czdz

dtdt (6)

177 whereT 5oscillation period.

178 In contrast, the power associated with the unperturbed ow is 179 Pf5rU3b

m=2, and therefore, the energy harvesting efficiency is 180 h5Pg=Pf .

181 The direct integration of Eq. (6), after the introduction of the 182 dimensionless variables, to obtainPgas performed in Barrero-Gil 183 et al. (2010), results in a value for the efficiency of

h¼1 2

a1 Ap Up 2 þ3 4a3 Ap Up 4 þ5 8a5

Ap Up 6 þ35 64a7 Ap Up 8 þ 63 128a9 Ap Up 10

þ/ (7)

184 A different approach can be introduced here for the determination 185 of the efficiency in a simpler way. The idea is to evaluate the power

186 P

g, not the aerodynamic force used in Eq.(6),fz, but the dissipative 187 term in Eq.(1), that is, the left-hand side of Eq.(1). Eq.(6)forPg 188 then becomes

Pg¼m

T ð T

0 d2z dt2

dz

dtdtþ2zvn ðT

0 dz dt

2 dtþv2n

ðT

0 zdz

dtdt 2 6 4 3 7 5

¼2mzvn

T ðT 0 dz dt 2

dt (8)

189 In writing the last term of Eq.(8), it has been considered that,zbeing 190 a periodic function, the integrals of both ðd2z=dt2Þ3ðdz=dtÞ and

191 z3ðdz=dtÞare zero in a cycle.

192 Then, because it has been assumed thatz5AðtÞcos½vnt1fðtÞ,

193 and the introduction of the dimensionless variables already

194 having been defined, gives c5ApðtÞcos½t1fðtÞ, and dz=dt

195

5vnbmdc=dt5 2vnbmApsinðt1fÞ, where it has been considered

196 that the time variation of bothApðtÞandfðtÞis so slow that they

197 can be considered constant in a cycle. Therefore

Pg¼z

mv3nb2m p ð 2p 0 dc dt 2 dt

¼zmv3nb2m p

ð 2p

0

f2ApðtÞsin½tþfðtÞg2dt¼zmv3nb2mAp2 (9)

198 Thus, the expression giving the efficiency now results in

h¼2zmp

Up Ap Up 2 (10) 199 wherezmpis proportional to the Scruton number. Obviously, this

200 last expression states that the only energy that is harvested out of

201 the flow by the galloping motion is the energy dissipated by

202 the structure.

203 The analytical calculation of the maximum efficiency becomes

204 too complex when more than thefirst two terms in the polynomial

205 approximation toczðaÞ, Eq.(3), are considered. In that case, the

206 problem solution is better achieved by a numerical method. The

207 procedure used in the present work when more than two terms are

208 used in Eq.(3)is the following:

209 1. First, the coefficientsa2je1of the series expansion in Eq.(3)are

210 calculated byfitting this expression to available experimental

211 data. Note that the values of thefitted coefficients depend on

212 the number of terms considered in Eq.(3).

213 2. Once the coefficients a2je1 are known, the corresponding

214 functions dAp=dt, Eq.(5), are evaluated. It must be pointed

215 out that the parameters involved in Eq.(5)areAp,Up, andzmp;

216 therefore, in this process, two of them arefixed (namely,Ap

217 andzmp), whereas the third one,Up, is considered to be an

218 independent variable. The result depends on the values chosen

219 forApandzmp(additional comments on this point are given

220 later in this work).

221 3. The roots of dAp=dt50 are then calculated; thus, the critical

(4)

223 zmpare obtained (formally, the smaller value of the criticalUp 224 is the one of interest).

225 4. Then, the efficiency is determined by using either Eq.(7)or 226 Eq.(10)as a function ofUp,Ap, andzmp.

227 Analysis of the Energy Harvesting Properties of 2285 Several Bluff Bodies

229 In the simplest approach, when only therst two terms (a1anda3) 2306 are considered in the series (3), the maximum efficiency occurs when

231 h

max5 2a21=ð6a3Þ. This expression gives a rough approximation to

232 the quality of a given two-dimensional body to extract energy 233 through transverse galloping oscillations. Despite the limited ap-234 plication of this approximation, it seems clear that the more 235 promising bodies for power extraction are those whose cross sec-236 tions provide large positive values of the coefcienta1 and small 237 negative values ofa3. In afirst approximation (small values of the 238 angle of attack), the coefficienta1can be identified with the Glauert-239 Den Hartog parameter for galloping, and it must be positive (the 240 body must be stalled), whereasa3has to be negative, because this 241 parameter represents the capacity of the system to keep the amplitude 242 of the galloping oscillations betweenfinite boundaries (limit cycle 243 oscillations).

244 Note also that when only two terms are considered, the maximum 245 efciency does not depend on the parameters involved in Eq.(10). 246 This expression is not applicable, however, when more terms are 247 retained in Eq.(5).

248 Using sufficient terms in the series expansion of the force co-249 efficientczðaÞis important for properly predicting the energy har-250 vesting efficiency of a given structure. In addition, when this force 251 coefficient is obtained from experimental data, the process offitting 2527 these data determines the accuracy of the results. Thetting process 253 is affected by the value of the reference anglea0selected [formally, 254 only the antisymmetric part ofczðaÞobtained from experimental 255 data are considered] and by the number of data points taken around 256 the central one. In thefitting process of a given equation to a set of 257 experimental data, thefirst condition that should be fulfilled is that 258 the equation must be appropriate to reproduce the main features of 259 the experimental data (in the case of a power series, this is generally 260 achieved by retaining enough terms in the truncated series), oth-261 erwise, very different conclusions can be reached from the con-262 clusions that the experimental data show.

263 The main purpose of this work is to investigate how the ap-264 proximation chosen to derive the force coefficient from the experi-265 mental data for calculating the energy harvesting efficiency 266 determines the accuracy of the results. The criteria derived for that 267 approximation, together with the simpler approach to evaluate the 268 harvesting efciency presented in theMathematical Model sec-269 tion, substantially improves the method presented in Barrero-Gil 270 et al. (2010). This improved method is applied to bodies with dif-271 ferent cross sections.

272 The analysis is performed first on D-shaped cross sections, 273 whose efficiency is derived in Barrero-Gil et al. (2010) by choosing 274 a cubic polynomial for the force coefficient approximation from 275 experimental data available in the literature. To contrast these results 276 and apply the new findings to the energy harvesting efficiency 277 evaluation, new wind tunnel tests were performed at IDR/UPM on 278 a D-shaped cross-section body.

279 A prism with the cross section (Fig. 2) was built up, and its 280 aerodynamic properties were measured in the A9 wind tunnel at 281 IDR/UPM. This wind tunnel is an open circuit, closed test chamber 282 whose cross section is 1.8-m high and 1.5-m wide. Turbulence 283 intensity is approximately 2.5% at 20 m=s wind velocity. The model 284 span was 0.8 m, and there were twoat plates parallel to theow at

285 the ends of the prism to assure two-dimensional behavior.

Aero-286 dynamic loads were measured through a six-component

strain-287 gauge balance (ATI DELTA, ATI Industrial Automation, Apex, 8

288 North Carolina), whereas theflow velocity was measured with a

289 Pitot tube (model 3.3.311, Airflow Systems, Dallas, Texas), connected

290 to a Schaewitz Lucas pressure transducer. The D-shaped model

291 was mounted on a rotating platform NEWPORT RV120-PP-HL

292 (Newport Corporation, Irvine, California), which, in turn, was

293 mounted on the strain-gauge balance. The rotating platform can set

294 the model angle of attack within 60:1°accuracy. The balance has

295 a maximum measurement uncertainty of 1.25%. No provisions for

296 blockage corrections of the measured loads have been undertaken, as

297 even in the worst case, the model front area (including end plates) is

298 less than 8% of the test sectionflow area.

299 Aerodynamic forces were measured from a05 290°toa0

300

5 190°at a 1°step. At each reference angle, the loads were sampled

301 at 200 Hz during 30 s. The results obtained are shown in Fig.2, in the

302 same plot of the experimental data reported by Barrero-Gil et al.

303 (2010). In Fig. 2, the symbols represent the experimental data

304 reported by Barrero-Gil et al. (2010) (circles) and the data measured

305 at IDR/UPM facilities (rhombi). The lines correspond to fitted

306 equations in the experimental data reported by Barrero-Gil et al.

307 (2010) (circles) by retaining only thefirst two terms in Eq.(3)(solid

308 lines). The influence of the number of experimental points in the

309 fitting process is as follows: line 1 represents thefitting result when

310 only nine experimental points (white circles) are used; line 2

rep-311 resents thefitting result when 11 experimental points (white and gray

312 circles) are considered, and line 3 is obtained when the whole set of

313 experimental points reported by Barrero-Gil et al. (2010) is

con-314 sidered in thefitting process [15 experimental points (white, gray,

315 and black circles)]. Line 4 is the one given in Barrero-Gil et al.

316 (2010),fitting with only two terms.

317 The dependence on the considered number of experimental

318 points used tofit lateral force coefficients to data presented in Fig.2

319 can be found in Table1.

320 There are some differences between the two sets of experimental

321 data, which can be explained by keeping in mind the experiments

322 described herein, that the noncircular part of the D-shaped body

(5)

323 contour is not exactly a flat surface, but is formed by two planes, 324 creating an angle of 170° between them (instead of 180°). This 325 slightly modifies the aerodynamic behavior at large values of the 326 angle of attack (in this case, the relative thickness is 1.84 instead of 2). 327 The force coefcient curve is then tted by taking different 328 numbers of experimental points measured during the IDR/UPM 329 wind tunnel tests to assess the inuence of this parameter, and 330 thus, thefitting curves using the 9, 11, and 15 experimental points 331 are also represented in Fig. 2. Finally, the tting curve given in 3329 Barrero-Gil et al. (2010) is also included in Fig.2. Aclose look at the 333 curves in Fig.2indicates how important it is to have the right choice 334 of the approximation toczðaÞfrom the experimental data; specifi-335 cally, if the range of experimental data points is large (this applies to 336 D-shaped cross sections but also to other bodies), then the curvature 337 variation of the function should not be so simple. As shown in Fig.2, 338 the real behavior of the experimental data is characterized by an 339 almost zero slope at the origin, but if the data interval taken for the 340 curvefitting is too large, the extreme points of the interval are those 341 driving the value of the slope of thefitting curve. In this case, it is 342 possible to obtain afitting curve with a slope at the origin larger than 343 that shown by the experimental data, which means an overestimation 344 of the energy harvesting efficiency. The curvature of the curves at the 345 origin is also changed.

346 The new approach for the evaluation of the energy harvesting 347 efficiency is also applied to other two-dimensional bodies that may 348 be unstable and experience transverse galloping oscillations. During 349 the past several years, a research program devoted to studying the 350 transverse galloping characteristics of prismatic bodies with simple 351 basic cross-sectional shapes has been carried out at IDR/UPM. 352 Within this program, parametric experimental studies of bodies 353 like isosceles triangles, ellipses, biconvex aerofoils, and rhombi

354 (Fig.3) have been carried out by wind tunnel tests. The parameters

355 that define the members of a given family of bluff bodies are the

356 relative thickness,t, which is defined as the ratio of the maximum

357 thickness,bm, of the reference body cross section to its chord,c. The

358 reference body cross section is the one corresponding toa050

359 (Fig.3). SimilarczðaÞcurves for these bodies, as the one depicted in

360 Fig.2, can be found in Alonso and Meseguer (2006), in Alonso et al.

361 (2005,2007) in the case of triangular cross section bodies, in Alonso

362 et al. (2009) for biconvex cross section bodies, in Alonso et al.

363 (2009) and in Ibarra et al. (2014) for rhomboidal cross-section

364 prisms, and in Alonso et al. (2010) for elliptical cross-section

365 beams. In all the preceding cases, the dependence ofczðaÞon the

366 parameters defining the section geometry is considered.

367 Obviously, to extract energy from galloping oscillations, the

368 configuration must develop instability, which according to

pub-369 lished information, only occurs at given values of the reference

370 anglea0, which is different for each family of bodies. The

differ-371 ent stability maps for the previously listed geometries have been

372 published elsewhere. Information concerning triangular prisms can

373 be found in Alonso et al. (2005,2007,2012), Alonso and Meseguer

374 (2006), and Iungo and Buresti (2009); elliptical bodies are

consid-375 ered in Alonso et al. (2010), and both biconvex and rhomboidal

376 cross-section bodies are analyzed in Alonso et al. (2009).

377 The corresponding values of the function czðaÞ5½cdðaÞtana

378

1clðaÞ=cosahave been calculated from the available wind tunnel

379 tests data on the aerodynamic coefficients of the bodies under

380 consideration. These data arefitted to polynomials in tana, thus

381 giving the values of the coefficients a2je1. Then, the maximum

382 efficiencies are obtained according to Eq.(10)and the new approach

383 presented in the“Mathematical Model”section.

384 The geometry that shows the maximum efficiency now needs to

385 be determined, as well as which geometry (represented by its relative

386 thickness) also shows the maximum efficiency for each family of

387 bodies. Fig.4shows the variation with the relative thickness of the

388 maximum efficiency of the different families of bluff bodies. The

389 results represented in Fig. 4 correspond to the case Ap52 and

390 zmp51ða0 ≅180°Þ, but, as has already been mentioned, these

391 results will be different if other values of the parameters concerned

392 are considered. Such behavior is shown in Fig.5, where the variation

393 of the maximum efficiency with the dimensionless amplitude Ap,

394 in the case of isosceles triangular bodies with main vertex angle

395 b530°, is shown. There are two families of curves in this plot.

396 Those that are almost vertical straight lines come from Eq.(7)for

397 the giving values ofUp, whereas the curved lines are obtained from

398 Eq.(10), assuming somefixed values ofzmp.

399 From this plot, it can be realized that onceUpisfixed, the

am-400 plitude of oscillationsApfor maximum efficiency isfixed, because

Fig. 3.Two-dimensional prism families whose energy harvesting suitability under galloping oscillations are considered: (a) triangle; (b) ellipse; (c) biconvex; (d) rhombus

Table 1.Variation of thea1anda3Coefficients Appearing in Eq.(3)of the Lateral Force Acting on D-shaped Bodies

17

Line number

Number of experimental

points a1

18 a3

1 11 20:159 20:758

2 9 20:305 20:394

3 15 20:601 0.043

4 The same as line 3, but

(0,0) excluded

20:79 0.19

(6)

401 the maximum position is obtained by differentiation of Eq.(7), and 402 zmpdoes not have an influence. The maximum efficiency grows as 403 the parameterzmpgrows [as deduced from Eq.(10)], and similar 404 conclusions can be obtained by considering any other order of the 405 parameters involved. The reference conguration (Ap52,zmp51)

406 used to calculate the maximum efficiency of the different bluff

407 bodies is represented by a circle in Fig.5.

408 By keeping in mind Eqs. (7) and(10), each geometrical

con-409 figuration can be represented by a single point if the self-similar

410 variablesAp=Up and hmaxUp=ð2zmpÞare used instead ofAp and

411 hmax. In these self-similar variables, Fig.5reduces toAp=Up50:77

412 andhmaxUp=ð2zmpÞ50:35.

413 The values of the self-similar variables change with the body

414 shape, but it is remarkable that, within the validity range of this

415 simplified model, the transverse galloping energy harvesting

prop-416 erties of a given two-dimensional body can be summarized in

417 a single point in theAp=UpversushmaxUp=ð2zmpÞ plane.

418 Concerning Fig.4, in the case of triangular cross-section prisms,

419 which are identified by the angle b of the main vertex, tanb

420

5bm=ð2cÞ, there are several instable regions in the a0-bplane

421 (Alonso et al. 2005,2007,2009,2010,2012;Alonso and Meseguer

422 2006;Iungo and Buresti 2009). One of these regions is in the vicinity

423 ofa05180°, whereas another one appears at approximately a0

424

560°. Obviously, to obtain energy from the galloping oscillations,

425 the body under consideration must undergo instability, which means

426 that at least the Glauert-Den Hartog criterion must be satisfied.

427 Taking this fact into consideration, only those configurations that

428 clearly meet this criterion (a0≅180°) have been considered in

429 Fig.4.

430 According to Fig.4, triangular cross-section bodies ata05180°

431 seem to be the more promising ones for energy harvesting from the

432 galloping phenomenon.

433 Conclusions

434 An improved approach to estimate the energy harvesting efficiency

435 from the transverse galloping of bluff bodies using experimental

436 data has been presented. The importance of properly choosing the

437 process offitting the polynomial curve of the aerodynamic force

438 coefficient to the experimental set of data has been demonstrated

439 with existing experimental data from bodies with different cross

440 sections and with data obtained from new experiments. It has been

441 proven that, unless the aerodynamic force coefficient experimental

442 curve is extremely simple, experimental data cannot befitted by

443 a cubic polynomial, and higher powered terms must be retained in

444 the series expansion (3)to reproduce the real behavior. 10

445 The inappropriate election of the polynomial used tofit the

ex-446 perimental data would give an unrealistic and an even too optimistic

447 capability of the given body to gather energy from windflow, or,

448 in contrast, lead to disregarding the given two-dimensional body

449 shape. This is based on a wrong estimation of the energy harvesting

450 capability, which seems to be the case of isosceles triangular

cross-451 section bodies in Barrero-Gil et al. (2010).

452 Another important conclusion is that, although there are many

453 cross-section shapes that show transverse galloping instability in

454 a range of angles of attacka0, from the point of view of energy

455 harvesting, only a few of them seem to be appropriate to give

456 reasonable values of the maximum efficiency associated with the

457 extraction of energy from the incident wind. Values of the maximum

458 efficiency of different families of prisms are provided and compared.

459 The most promising configuration seems to correspond to the

re-460 verse triangular cross-section prisms, and if the study is confined

461 to isosceles triangles (identified by the anglebof the main vertex),

462 the maximum efficiency is reached for triangles with a vertex angle

463 close tob560°.

464 Further experimental investigations concerning triangular bodies

465 (Ibarra et al. 2014) demonstrated that inverted isosceles triangles

466 develop galloping instabilities atUp≅ 35; in the case of isosceles

Fig. 4.Maximum efficiency,hmax, versus relative thickness,t, of the

different families of prisms shown in Fig. 3; triangles at a0≅180°

(triangles); rhombi (rhombi), biconvex (squares), and ellipses (circles); solid circles represent results of D-shaped bodies; vertical bars indicate the estimated error

15

Fig. 5. Variation of the maximum efficiency for energy harvesting, hmax, with the dimensionless amplitude of galloping oscillations,Ap, of a

two-dimensional bluff body whose cross section is an isosceles triangle with a main vertexb530°; results correspond toa05180°; numbers

on the curves indicate the values of the reduced velocity,Up(vertical lines) or the product of the reduced mass and the damping ratio,zmp; the circle (Ap52,zmp51) represents the reference configuration selected to compare the efficiency of different galloping bluff bodies (Fig.4

(7)

467 triangles with a main vertexb560°, whereUpstands for the re-468 duced velocity,Up5U=ðvncÞ, whereUis the air speed,vnis the 469 first natural body frequency, andcis a characteristic length. Before 470 galloping occurs, at lower wind velocities, Up3, a vortex-induced 471 vibration (VIV) episode appears. However, this VIV episode is of 472 less interest from the point of view of wind energy harvesting, 473 because VIV takes place at low air velocity (thus, the wind energy to 474 be gathered is small) and because of the small amplitude of the body 475 oscillations.

476 In galloping oscillations, these two effects are enlarged. In the 477 experiments reported by Ibarra et al. (2014), the galloping phe-478 nomenon appears forUp.35, whereas VIV oscillations occur at 479 Up≅3, as already stated, which implies the galloping wind kinetic 480 energy is more than 100 times greater than the one associated with 481 vortex shedding. The oscillation amplitudes are highly dissimilar, 482 and the ratio between galloping amplitude oscillation, AG, and 483 vortex-induced one, AV, can be up to AG=AV$4. In galloping 484 experimentation, the allowed maximum oscillation amplitudes must 485 be limited, and the experimental sequences should be stopped when 486 such maxima of the oscillation amplitudes are reached, to avoid any 487 damage to the setup.

488 From a practical point of view, the advantages of energy har-489 vesting by using inverted isosceles triangles are probably limited 490 by the drawbacks of large amplitude oscillations and the potential 491 destructive effects of galloping instabilities, but the aspects of these 492 problems are out of the scope of this paper.

493 References

494 Abdelkefi, A., Hajj, M. R., and Nayfeh, A. H. (2012a).“Power harvesting

495 from transverse galloping of square cylinder.”Nonlinear Dyn., 70(2),

496 1355–1363.

497 Abdelkefi, A., Hajj, M. R., and Nayfeh, A. H. (2012b).“Phenomena and

498 modeling of piezoelectric energy harvesting from freely oscillating

499 cylinders.”Nonlinear Dyn., 70(2), 1377–1388.

500 Abdelkefi, A., Hajj, M. R., and Nayfeh, A. H. (2012c).“Modeling and

501 analysis of piezoaeroelastic energy harvesters.”Nonlinear Dyn., 67(2),

502 925–939.

503 Abdelkefi, A., Hajj, M. R., and Nayfeh, A. H., (2013a).“Piezoelectric energy

504 harvesting from transverse galloping of bluff bodies.”Smart Mater. 505 Struct., 22(1), 015014.

506 Abdelkefi, A., Nayfeh, A. H., and Hajj, M. R. (2012d).“Design of

pie-507 zoaeroelastic energy harvesters.”Nonlinear Dyn., 68(4), 519–530.

508 Abdelkefi, A., and Nuhait, A. O. (2013).“Modeling and performance

509 analysis of cambered wing-based piezoaeroelastic energy harvesters.”

510 Smart Mater. Struct., 22(9), 095029,

511 Abdelkefi, A., Scanlon, J. M., McDowell, E., and Hajj, M. R. (2013b).

512 Performance enhancement of piezoelectric energy harvesters from

513 wake galloping.”Appl. Phys. Lett., 103, 033903.

514 Abdelkefi, A., Yan, Z., and Hajj, M. R. (2013c).“Modeling and nonlinear

515 analysis of piezoelectric energy harvesting from transverse galloping.”

516 Smart Mater. Struct., 22(2), 025016.

517 Abdelkefi, A., Yan, Z., and Hajj, M. R. (2014).“Performance analysis of

518 galloping-based piezoaeroelastic energy harvesters with different

cross-519 section geometries.”J. Intell. Mater. Syst. Struct., 25(2), 246–256.

520 Akaydin, H. D., Elvin, N., and Andreopoulos, Y. (2010).“Wake of a

cyl-521 inder: A paradigm for energy harvesting with piezoelectric materials.”

522 Exp. Fluids, 49(1), 291–304.

523 Alonso, G., and Meseguer, J. (2006).“A parametric study of the galloping

524 instability of two-dimensional triangular cross-section bodies.”J. Wind 525 Eng. Ind. Aerodyn., 94(4), 241–253.

526 Alonso, G., and Meseguer, J. (2014). “Hysteresis phenomena in

aero-527 dynamics.”Hysteresis: Types, applications and behavior patterns in com-528 plex systems, J. C. Dias, ed., Nova Science Publishers, Hauppauge, NY.

529 Alonso, G., Meseguer, J., and Pérez-Grande, I. (2005).“Galloping

oscil-530 lations of two-dimensional triangular cross-sectional bodies.” Exp. 531 Fluids, 38(6), 789–795.

532

Alonso, G., Meseguer, J., and Pérez-Grande, I. (2007).“Galloping stability

533

of triangular cross-sectional bodies: A systematic approach.”J. Wind

534 Eng. Ind. Aerodyn., 95(9–11), 928–940.

535

Alonso, G., Meseguer, J., Sanz-Andrés, A., and Valero, E. (2010).“On the

536

galloping instability of two-dimensional bodies having elliptical

cross-537

sections.”J. Wind Eng. Ind. Aerodyn., 98(8–9), 438–448.

538

Alonso, G., Sanz-Lobera, A., and Meseguer, J. (2012).“Hysteresis

phe-539

nomena in transverse galloping of triangular cross-section bodies.”

540 J. Fluids Struct., 33, 243–251.

541

Alonso, G., Valero, E., and Meseguer, J. (2009).“An analysis on the

542

dependence on cross section geometry of galloping stability of

two-543

dimensional bodies having either biconvex or rhomboidal cross

sec-544

tions.”Eur. J. Mech. B/Fluids, 28(2), 328–334.

545

Bae, J. S., and Inman, D. J. (2014).“Aeroelastic characteristics of linear

546

and nonlinear piezo-aeroelastic energy harvester.”J. Intell. Mater. Syst.

547 Struct., 25(4), 401–416.

548

Barrero-Gil, A., Alonso, G., and Sanz-Andrés, A. (2010).“Energy

har-549

vesting from transverse galloping.”J. Sound Vibrat., 329(14), 2873–

550

2883.

551

Bryant, M., Wolff, E., and Garcia, E. (2011).“Aeroelasticflutter energy

552

harvester design: The sensitivity of the driving instability to system

553

parameters.”Smart Mater. Struct., 20(12), 125017.

554

Den Hartog, J. P. (1965).Mechanical vibrations, 4th Ed., McGraw Hill,

555

New York.

556

Doaré, A., and Michelin, S. (2011). “Piezoelectric coupling in

energy-557

harvestingflutteringflexible plates: Linear stability analysis and

con-558

version efficiency.”J. Fluids Struct., 27(8), 1357–1375.

559

Erturk, A., Vieira, W. G. R., De Marqui C., Jr, and Inman, D. J. (2010).“On

560

the energy harvesting potential of piezoaeroelastic systems.”Appl. Phys.

561 Lett., 96, 184103.

562

Glauert, H. (1919). “The rotation of an aerofoil about a fixed axis.”

563 (British) Advisory Committee for Aeronautics (ARC) R & M No. 595,

564 Tech. Rep. of ARC for 1918–1919, His Majesty’s Stationery Office,

565

London, 443–447.

566

Grouthier, C., Michelin, S., and de Langre, E. (2012).“Energy harvesting

567

using vortex-induced vibrations of tensioned cables.” 11

568

Hobbs, W. B., and Hu, D. L. (2012).“Tree-inspired piezoelectric energy

569

harvesting.”J. Fluids Structures, 28, 103–114.

570

Ibarra, D., Sorribes, F., Alonso, G., and Meseguer, J. (2014).“Transverse

571

galloping of two-dimensional bodies having a rhombic cross-section.”

572 J. Sound Vibrat., 333(13), 2855–2865.

573

Iungo, G. V., and Buresti, G. (2009).“Experimental investigation on the

574

aerodynamic loads and wakeflow features of low aspect-ratio triangular

575

prisms at different wind directions.”J. Fluids Struct., 25(7), 1119–

576

1135.

577

Jung, H.-J., and Lee, S. W. (2011).“The experimental validation of a new

578

energy harvesting system based on the wake galloping phenomenon.”

579 Smart Mater. Struct., 20(5), 055022.

580

Jung, H.-J., Lee, S. W., and Jang, D. D. (2009).“Feasibility study on a

581

new energy harvesting electromagnetic device using aerodynamic

in-582

stability.”IEEE Trans. Magn., 45(10), 4376–4379.

583

Luo, S. C., Chew, Y. T., Lee, T. S., and Yazdani, M. G. (1998).“Stability

584

to translational galloping vibration of cylinders at different mean angles

585

of attack.”J. Sound Vibrat., 215(5), 1183–1194.

586

Mackowski, A. W., and Williamson, C. H. K. (2013).“An experimental

587

investigation of vortex-induced vibration with nonlinear restoring forces.”

588 Phys. Fluids, 25, 087101.

589

Mehmood, A., Abdelkefi, A., Hajj, M. R., Nayfeh, A. H., Akhtar, I., and

590

Nuhait, A. O. (2013).“Piezoelectric energy harvesting from

vortex-591

induced vibrations of circular cylinder.”J. Sound Vibrat., 332(19),

592

4656–4667.

593

Naudascher, E., and Rockwell, D. (1994).Flow induced vibrations—An

594 engineering guide, A. A. Balkema, Rotterdam, Netherlands.

595

Ng, Y. T., Luo, S. C., and Chew, Y. T. (2005).“On using high-order

596

polynomial curve fits in the quasi-steady theory for square-cylinder

597

galloping.”J. Fluids Struct., 20(1), 141–146.

598

Novak, M. (1969).“Aeroelastic galloping of prismatic bodies.”J. Engrg.

599 Mech. Div., 96, 115–142.

600

Novak, M. (1972).“Galloping oscillations of prismatic structures.”J. Engrg.

(8)

602 Novak, M., and Tanaka, H. (1974). “Effect of turbulence on galloping

603 instability.”J. Engrg. Mech. Div., 100, 27–47.

60420 Païdoussis, M., Price, S., and de Langre, E. (2011). Fluid–structure 605 interactions: Cross-flow-induced instabilities, Cambridge University

606 Press, Cambridge, U.K., 15–104.

607 Peng, J., and Chen, G. S. (2012).“Flow-oscillating structure interactions and

60812 the applications to propulsion and energy harvest.”Appl. Phys. Res., 4(2).

609 Peng, Z., and Zhu, Q. (2009). “Energy harvesting throughflow-induced

610 oscillations of a foil.”Phys. Fluids, 21, 123602.

611 Sewatkar, C. M., Sharma, A., and Agrawal, A. (2012).“On energy transfer

612 inflow around a row of transversely oscillating square cylinders at low

613 Reynolds number.”J. Fluids Struct., 31, 1–17.

614 Sirohi, J., and Mahadik, R. (2011).“Piezoelectric wind energy harvester for

615 low-power sensors.”J. Intell. Mater. Syst. Struct., 22(18), 2215–2228.

616 Sirohi, J., and Mahadik, R. (2012).“Harvesting wind energy using a

gal-617 loping piezoelectric beam.”J. Vib. Acoust., 134, 011009.

618

St. Clair, D., Bibo, A., Sennakesavababu, V. R., Daqaq, M. F., and Li, G.

619

(2010).“A scalable concept for micropower generation using fl

ow-620

induced self-excited oscillations.”Appl. Phys. Lett., 96, 144103.

621

Tang, L., Païdoussis, P., and Jang, J. (2009).“Cantileveredflexible plates in

622

axialflow: Energy transfer and the concept offlutter-mill.”J. Sound

623 Vibrat., 326(1–2), 263–276.

624

Wu, X., Ge, F., and Hong, Y. (2012).“A review of recent studies on

vortex-625

induced vibrations of long slender cylinders.”J. Fluids Struct., 28, 292–308. 13

626

Zhao, L., Tang, L., and Yang, Y. 2012.“Small wind energy harvesting from

627

galloping using piezoelectric materials.”ASME 2012 Conf. on Smart

628 Materials, Adaptive Structures and Intelligent Systems, ASME, New

629

York, 919–927.

630

Zhu, Q. (2011).“Optimal frequency forflow energy harvesting of aflapping

631

foil.”J. Fluid Mech., 675, 495–517.

632

Zhu, Q., Haase, M., and Wu, C. H. (2009).“Modeling the capacity of a

633

(9)

Q:A_NEW! ASCE Open Access: Authors may choose to publish their papers through ASCE Open Access, making the paper freely available to all readers via the ASCE Library website. ASCE Open Access papers will be published under the Creative Commons—Attribution Only (CC-BY) License. The fee for this service is $1750, and must be paid prior to publication. If you indicate Yes, you will receive a follow-up message with payment instructions. If you indicate No, your paper will be published in the typical subscribed-access section of the Journal.

Q:1_Please check that ASCE Membership Grades (Member ASCE, Fellow ASCE, etc.) are provided for all authors who are members.

Q:2_Per journal style, please provide the complete names for ‘‘IDR/UPM’’ and ‘‘E.T.S.I.’’ (These will also be expanded at first instance in the article, per style.)

Q:3_In the equation ‘‘d(a). . .,’’ the ‘‘d’’ isitalic, while the ‘‘d’s’’ used previously are roman; please confirm, throughout the article, whether ‘‘d’’ should instead be ‘‘d’’.

Q:4_In Eq. (4) and subsequent equations, the asterisk (*) has been set as a superscript; please confirm or correct.

Q:5_Per journal style, there may not be a single subhead in a section; please add at least one subhead to this ‘‘Mathematical Model’’ section, or delete this subhead.

Q:6_Please clarify if this (3) refers to Eq. (3) or to the numbered list above this section.

Q:7_The sentence beginning ‘‘The fitting process is affected...’’ was edited for readability; please confirm that it retains intended meaning.

Q:8_Per style, all materials and equipment mentioned in the article must be identified by the manufacturer’s name and location; please confirm/ correct ATI DELTA.

Q:9_Are changes to the sentence ‘‘A close look at the curves...’’ okay?

Q:10_Please clarify if ‘‘(3)’’ refers to point 3 in the numbered list or to Equation (3).

Q:11_For Grouthier et al. (2012), please confirm whether this article was published in print, as well as online; if in print, please provide complete citation.

Q:12_For Peng and Chen (2012), please provide complete page range.

Q:13_Reference ‘‘Wu, Ge, Hong, 2012’’ is not cited in the text. Please add an in-text citation or delete the reference.

Q:14_In Fig. 1, the list of variables has been alphabetized per journal style; please confirm/correct.

Q:15_Because Fig. 4 will print in black and white; ‘‘black circles’’ has been changed to ‘‘solid circles’’ to differentiate from the green-outlined circles—green-outlined circles will print in shades of gray.

Q:16_Please add brief explanation stating how Fig. 4 is related to Fig. 5, as was done in Fig. 4 (i.e., ‘‘prisms shown in Fig. 3’’).

Q:17_The original Table title for Table 1 was too long; the text has been added as a table footnote; please confirm that this does not alter intended meaning.

Q:18_Column header changed from ‘‘A1’’ to ‘‘a1,’’ per table title; please confirm or correct.

Q:19_Paı¨doussis et al. 2010 is not included in the reference list; please correct or delete the in-text citation.

Q:20_Paı¨doussis et al. 2011 is not mentioned in the article text; please correct or delete this reference.

AUTHOR QUERIES

Referencias

Documento similar

Therefore, the current study introduces the optimal configuration of renewable energy generation systems for Chiang Mai University, which is one of the largest public universities

Given the much higher efficiencies for solar H 2 -generation from water achieved at tandem PEC/PV devices ( > 10% solar-to-H 2 energy efficiency under simulated sunlight) compared

Figure 5.25: (Lower panel) Vibrationally resolved cross section for the three first vibrational levels of the 1σ −1 photoionization of CO as a function of the photon energy..

This corresponds to 55.57MJ Equivalent Primary Energy (EPE) assuming a 35% of efficiency of primary energy to electrical energy conversion as seen in Table 4.28,

Our contributions are as follows: i) we present an accurate analytical model that is able to predict the energy consumption, ii) we present an approximate model that sacrifices

From the figure is obvious to see that, while the throughput maximization provides the same CW for a given number of stations, the optimal CW for energy efficiency depends

The key idea underlying the energy efficiency of network sharing is that network capacity supply can be modulated by switching off some networks for the time periods in which traffic

The analysis presented shows that energy efficient motors are an opportunity for improving the efficiency of motor systems, leading to large cost-effective energy savings,