Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Accepted Manuscript

Title: Dynamic Failure of Clamped Metallic Circular Plates Subjected to


Underwater Impulsive loads

Author: Wei Huang, Bin Jia, Wei Zhang, Xianglin Huang, Dacheng Li, Peng
Ren

PII: S0734-743X(16)30195-6
DOI: http://dx.doi.org/doi: 10.1016/j.ijimpeng.2016.04.006
Reference: IE 2679

To appear in: International Journal of Impact Engineering

Received date: 22-12-2015


Revised date: 15-4-2016
Accepted date: 17-4-2016

Please cite this article as: Wei Huang, Bin Jia, Wei Zhang, Xianglin Huang, Dacheng Li, Peng
Ren, Dynamic Failure of Clamped Metallic Circular Plates Subjected to Underwater Impulsive
loads, International Journal of Impact Engineering (2016), http://dx.doi.org/doi:
10.1016/j.ijimpeng.2016.04.006.

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

2 Underwater Impulsive loads

3 Wei Huanga,b, Bin Jiaa, Wei Zhanga*, Xianglin Huanga, Dacheng Li, Peng Rena,c
a
4 Hypervelocity Impact Research Center, Harbin Institute of Technology, Harbin 150080, P.R. China
b
5 Department of Mechanical Engineering, Michigan State University, East Lansing, MI 48824,USA
c
6 School of naval architecture & ocean engineering, Jiangsu university of science and technology,

7 Jiangsu, P.R.China

9 Highlights:

10  Quantify the failure of solid plates to water-based impulsive has been

11 conducted;

12  Analyses focus on thickness, FSI parameter, and patch size of monolithic

13 plates;

14  Comparisons are conducted between water-backed and air-backed conditions;

15  Scalding relations are developed considering different effects;

16  Parameters for constitutive model of the material have been provided.

17

18 Abstract: The dynamic response and failure of monolithic metallic plates subjected to

19 water-based impulsive loads are investigated experimentally. The analysis focuses on

20 the effects of plate thickness, fluid-structure interaction parameter, and patch size of

21 loading area on deformation and failure modes in clamped solid 5A06 aluminum alloy

*
Corresponding author. Tel.: +86 451 86417978 13; fax: +86 451 86402055.

E-mail addresses: zhdawei@hit.edu.cn (W Zhang), 13b918024@hit.edu.cn(W Huang)

Page 1 of 32
22 plates under air-backed and water-backed loading conditions. The plates are subjected

23 to impulsive loads of different intensities using a projectile-impact based underwater

24 non-contact explosive simulator. 3D digital imaging correlation method is used to

25 capture the dynamic response of plates to make comparison with postmortem analysis.

26 Depending on the loading rate, the inelastic deformation is the primary failure mode

27 of the plates. The different linear relationships between deflection resistance and

28 applied impulse are identified experimentally, considering the influences of the effects

29 of plate thickness, fluid-structure interaction parameter, and patch size of loading area.

30 The results shows that the effect of loading area is the most influential factor on

31 transverse deflection. The results affirm that the plate under water-backed condition

32 shows a 53% reduction in the maximum plate deflection compared with the plate

33 under air-backed condition. Quantitative structure-load-performance relation is carried

34 out to facilitate the advanced study on metallic structures and provides guidance for

35 structural design.

36 Keywords: dynamic failure, fluid-structure interactions, blast resistance, monolithic

37 plate, experimental analysis

38 1. Introduction

39 Military and civilian ship structures, such as the hull and keel structures, are

40 exposed to various environmental loadings which include high and low temperature

41 extremes, transient impulsive loads, and corrosive sea water. Additionally, the

42 structures are designed to survive from both surface and underwater explosions and

43 weapons impact. The material properties, blast-resistant performances, and geometric

44 design of sub-structures must be well-understood and quantified.

45 Clamped structures are representative of the underwater vessel,which have


2

Page 2 of 32
46 attracted a great amount of interest to investigate the dynamic responses. Recently,

47 experimental and theoretical studies on the metallic and composite sandwich plates

48 have been conducted by many researchers [1-14]. Metallic solid and sandwich

49 structures have been studied in terms of constituent material behavior, structural

50 hierarchy, topological characteristics and complex loading involving fluid structure

51 interactions. Using light gas gun-based impact loading to generate exponentially

52 underwater pressure impulses, Deshpande [2]and Espinosa [15] designed novel

53 non-explosion impulsive simulators to exert planer pressure wave on targets. A

54 number of topological cores were investigated, including the corrugated core,

55 prismatic diamond core, honeycomb core, and metal foam [4, 8, 9, 16, 17].

56 Constitutive relations have been developed for sandwich structures, accounting for

57 dynamic crush behavior of core and plasticity in constituents by Deshpande [18, 19]

58 and Xue [20]. The deformation of sandwich is divided into three phases: phase I is the

59 fluid-structure interaction which is up to the point of the first cavitation of the fluid;

60 stage II is the core compression until the front and back faces get an equal velocity,

61 and is followed by the bending and stretching of stage III. McShane et al.[16]

62 analyzed the three phases by making comparisons among the fully decoupled model,

63 cross-coupled model, and fully-coupled model, to investigate the fluid-structure

64 interaction effect during the deformation of plates. The results indicate that the

65 Taylor’s analysis based on a free-standing front face-sheet underestimated the

66 transmitted momentum by 20–30% due to the continued fluid loading during the

67 whole deformation of sandwich. Fleck and Deshpande [19, 21] examined the

68 fully-coupled fluid-structure interaction in their analytical models to predict the

69 transmitted momentum and deformation of the metallic sandwich. The parallel

70 research [12] on different sandwich cores was based on the similar analytical model.
3

Page 3 of 32
71 These articles concluded that metallic sandwich structures outperform monolithic

72 plates when the deformation is dominated by bending. However, Schiffer [5, 17]

73 reported that sandwich plates may or may not outperform rigid plates of equal mass in

74 terms of the impulse imparted to the structure in a blast event.

75 The solid metallic panels are the basic components of significant metallic

76 structures, which have been studied experimentally and theoretically for several

77 decades. Neglecting the elastic effect, Jones [22] studied the rectangular and circular

78 solid metallic panels under different loading conditions and proposed the ‘bound’

79 solution for the structural dynamic response. Considering the bending and shearing,

80 Schiffer [7] developed a model for elastic deformation of composite solid plates

81 subjected to underwater impulsive loads, considering the effect of fluid-structure

82 interaction. Nurick [23] conducted comprehensive experiments on fully clamped

83 circular and rectangular steel plates subjected to blast loads. With the increase of

84 impulsive intensities, failure modes are divided into three phases: mode I, inelastic

85 deformation, which is caused by plastically bending and stretching; mode II, tearing at

86 the supports. Plate stretching is followed by tensile rupture at the supports; mode III,

87 shearing at supports. Shear failure occurs at the supports with negligible plastic

88 deformation in the remainder of the beam. The typical discing and petalling failure

89 modes in impulsively loaded clamped plates were analyzed by Lee and Wierzbicki

90 [24, 25], and the tensile tearing modes were reminiscent mode II of failure modes for

91 impulsively loaded beams. Balden [26] made experimental and numerical

92 investigations into the shear rupture modes (mode III) of impulsively loaded clamped

93 circular plates. Kazemahvazi [27] presented an experimental study on the failure

94 modes of low strength copper plates subjected to underwater blast loads. It was

95 concluded from the micrographs that the local failure mechanism is tensile necking,
4

Page 4 of 32
96 regardless of whether the macroscopic mode is petalling or shear-off. Zamani [28]

97 presented the results of analytical and experimental studies on the response of steel

98 and aluminum circular plates in two different media of air and water. A verified

99 empirical prediction of normal deflection was presented considering the material

100 strength, normal deflection, and intensity of impulses. Until now, detailed

101 experimental validation needs to be attempted to provide more correct analysis and

102 numerical predictions, especially in dynamic underwater situations for which the

103 literature is scarce.

104 The high strength-weight ratios and high stiffness-weight ratios are the

105 remarkable requirements of ship structures to resist transverse impulsive loads. Light

106 structures, such as sandwich and some aluminum alloy solid panels, outperform the

107 traditional steel plates in terms of these mechanical properties. To investigate the blast

108 resistance and failure modes of the clamped plate as a function of applied impulses,

109 the geometric property and loading configuration is important to the optimal design of

110 vessel structures [4]. Despite recent advances in understanding the dynamic response

111 of solid metallic plates, several issues remain. One is the lack of design relations that

112 quantify the dynamic response as functions of both geometric parameters and load

113 configurations. To obtain such relations, diagnostics which can provide in-situ,

114 time-resolved measurement are required to record the dynamic deformation.

115 Additionally, studies focused on plates which were in contact with water only on one

116 side and with air on the opposite side, but the plates in contact with water on both

117 sides were not considered especially in the experimental studies. The water-backed

118 plates is a more common condition for most marine structures.

119 The objective of this work is to identify the dynamic response of solid metallic

120 panels subjected to underwater impulsive loads experimentally. The focuses of present
5

Page 5 of 32
121 analysis are on understanding the deformation, failure modes and associated

122 mechanisms, and quantifying the blast resistance of panels as functions of plate

123 thickness, fluid-structure interaction parameter, and loading conditions. Experiments

124 are conducted under three distinct loading conditions: (1) an air-backed condition,

125 with the plate in contact with water on the impulse side, (2) a water-backed condition,

126 with both sides of the plate in contact with water, and (3) the plates subjected to

127 impulsive loading over a central loading patch, with the loading patch size r=0.7. The

128 results are presented in normalized forms to gain insight into underlying trends that

129 can be used to design more blast-resistant structures.

130 2. Fluid-structure interaction experiments

131 2.1 Experimental detail

132 In order to generate predictable and controlled high-intensity underwater

133 impulsive loads for testing marine structures, a projectile-impact based fluid–structure

134 interaction experimental simulator was designed to measure temporal and spatial

135 evolution and failure of structures, as shown in Fig. 1. A Planar pressure pulse is

136 generated by firing a projectile at a sliding piston. In order to obtain much higher

137 intensity of underwater impulses, the dimensions of the water chamber similar to that

138 used by Zhou [10] and Deshpande [2]. Important features of this setup include the

139 ability to generate pressure waves of a wide range of intensities, the ability to simulate

140 the loading of air-backed, water-backed, changeable loading areas and integrate

141 high-speed photography.

142 Fig. 1 shows the fully edge clamped plates under air-backed and water-backed

143 conditions respectively. The shock tube is a 500 mm long cylinder which is

Page 6 of 32
144 horizontally mounted and filled with water. It is made of armor steel and has an inside

145 diameter of 66 mm. A thin piston plate is mounted at the front end and the specimen is

146 located at the rear end. A projectile is accelerated by the gas gun and strikes the piston

147 plate, generating a planar pressure pulse in the shock tube. According to the analysis

148 of Deshpande [2], the mass of the projectile is an important factor affecting the peak

149 pressure and decay time of planer impulse. In order to obtain two different decay

150 times of the impulses, 5mm-thick (0.13kg) and 12mm-thick (0.22kg) projectiles are

151 used respectively. Initial velocities of projectile in the range of 20~220m/s are used to

152 delineate the effect of loading rate on the deformation of the structures. This velocity

153 range corresponds to peak pressures between 10~300MPa which are captured by the

154 pressure transducers [29] mounted at the top of water tube. Monolithic aluminum

155 plates of thickness 0.5mm, 1.0mm, 1.5mm are used.

156 Air-backed condition: Clamped specimens are tested by using the air-backed

157 shock simulator sketched in Fig. 1 (a). Six equally spaced clearance holes for bolts are

158 drilled into the aluminum plates on a pitch circle of radium 65mm, to clamp the plate

159 onto the end of the water tube. Two 0.5mm-thick annular rubber rings and a

160 5mm-thick annular steel ring are used to ensure that the specimens were edge

161 clamped fully. The effective loading region of clamped air-backed plates has the same

162 radius as the water column in this configuration, R=33mm. The 3D digital imaging

163 correlation method (DIC) is used to capture the dynamic responses of monolithic

164 plates temporally and spatially. Two high-speed cameras Phantom v12.2 are put at the

165 back of the specimen directly, in appropriate degrees, to ensure the error analysis

166 conducted in the calibration is acceptable and obtains accuracy and stability results. In

167 all of the tests, the cameras are ~25º from the axial line of the specimens. The selected

168 frame rate and resolution are 33,000 frames and 384×384 pixels respectively. During
7

Page 7 of 32
169 the calibration and post analysis, the business analysis software ARAMIS is used to

170 achieve the corresponding parameters.

171 Water-backed condition: Unlike the air-backed loading test, both sides of the

172 tested specimens are in contact with water, as shown in Fig. 1 (b). The length of back

173 water column is 150mm. The similar mounting method is used to clamp the plates

174 between the two water tubes which have identical inner radius. Two 0.5mm-thick

175 annular rubber rings with 66mm inner diameter are put at both sides of the monolithic

176 plate to ensure a good seal. The invisibility of the specimens’ dynamic deformation in

177 this condition results in that dynamic response cannot be captured by cameras.

178 Central loading condition: In this case, the plates are subjected to the air-backed

179 impulsive loading over a central loading patch, as shown in Fig. 1 (e). To obtain the

180 loading patch size 0.7, the inner diameter of steel annular ring on the back of

181 specimens is changed from 66mm to 96mm. Meanwhile, the loading area provided by

182 water remains 66mm as presented in the air-backed condition. Additionally, the

183 selected frame rate and resolution are changed to 20,000 frames and 512×512 pixels

184 respectively.

185 2.2 Summary of the fluid-structure interaction impulsive loads

186 As noted in [19, 21, 30, 31], the classical Taylor model [32] is entirely

187 satisfactory for monolithic plates. According to the analysis of one-dimensional

188 underwater blast waves for a plane wave impacting on a sliding plate, the impulse in

189 the fluid follows the relation

 t /
p  p0e (1)

190 where p 0 is the peak pressure of the impulse,  is the decay time. According to the

191 one-dimensional wave theory revealed by Deshpande [2], the peak pressure and decay
8

Page 8 of 32
192 time are adjusted independently by varying the velocity and mass of projectile

193 respectively. And the relation can be described in the form

m p (2
p0  cw  wv0 ,  
 wcw )

194 where cw ,  w are the speed of sound in water and density of water, v0 is the initial

195 velocity of the projectile, mp is the areal mass of the projectile. Regarding the full

196 reflection of the shock wave, the impulse applied on the stationary plate can be

197 calculated as

 t /
I0  2  p0e d t  2 p 0 (3)
0

198 The effect of fluid-structure interaction is regarded as a parameter  and the

199 initial time of cavitation is  c . The relation is given by

c 1
 ln  (4)
  1
200 where    w c w / m f
. The momentum per unit area I tr a n s transmitted into the plate

201 up to the instant of cavitation is given by

I tr a n s  I 0  (5)

1 
202 where   .

203 Parameter analysis dictates that the problem under investigation is governed by

204 the following set of independent non-dimensional parameters [27, 33],

I0  h
I  ,  ,h  (6)
 wcw A R R
205 where A is area of loading, R is radius of loaded region, δ is the normal deflection of

206 back side of the clamped plate, and h is thickness of target.

207 2.3 Material properties

208 Quasi-static tension and high temperature quasi-static tension tests are conducted
9

Page 9 of 32
209 with the plate material. Considering the high strain effect during the impulsive tests,

210 the effect of strain rate is performed for 5A06 aluminum alloy with Hopkinson bars.

211 Fig. 2 shows the geometry for quasi-static and dynamic tensile specimens.

212 These tests provide typical engineering stress-strain curves for different test

213 configurations, as shown in Fig. 3. In Fig. 3(a)~(b), the results are obtained from

214 standard quasi-static tensile tests on a smooth symmetric specimen at room temperature

215 and high temperatures at a nominal strain rate of 1.25×10-3/s on an Instron 5569

216 universal testing machine. As shown in Fig. 3 (a), the specimens are cut from the same

217 5A06 panel in different degrees to ensure the accuracy of tests. Additionally, an

218 extensometer with 25 mm gauge length is used to measure the engineering strain at

219 room temperature. The slight vibration of the curves in room temperature is caused by

220 the clamp of the machine. It is worth noting that 5A06 material presents an obvious

221 temperature softening effect. Fig. 3(c) shows the effect of strain rate hardening of the

222 metallic material using the SHTB. No obvious yield point is found for all of the

223 stress-strain curves, so the flow stress at 0.2% plastic strain is selected as the yield

224 strength for the material.

225 Johnson-Cook constitutive model for aluminum panel: Base on the tests

226 conducted on the materials, the materials herein are identified with obvious strain

227 hardening and high ductility. The Johnson-Cook constitutive model [34], which

228 accounts for the strain-hardening, temperature softening, and strain rate dependence, is

229 used to describe the metallic material’s mechanical response. The equation is described

230 as follow:

(7

  ( A  B  e q ) (1  C ln  e q ) (1  T * )
n * m
eq )

231 where A,B,C,m,n are model parameters,  eq


is Mises equivalent stress,  eq is the
10

Page 10 of 32
   

 eq   / 
*
232 equivalent plastic strain, eq 0 is nondimensional strain rate, eq is equivalent

233 plastic strain rate,  0 is reference strain rate, T *   T  Tr   Tm  Tr  is nondimensional

234 temperature and T, Tr, Tm are current temperature, conference room temperature and

235 melt temperature of 5A06 aluminum respectively. In order to use the Johnson-Cook

236 constitutive model to characterize the aluminum alloys in the further numerical

237 studies, the mechanical parameters are achieved as the fitted Johnson-Cook

238 parameters, as shown in Table 1.

239 3. Experimental results

240 A parametric study was carried out, focusing on the effects of (i) loads intensity,

241 (ii) thickness of plate, (iii) fluid-structure interaction parameter, (iv)patch size of

242 central loading area, and (v) air-backed and water-backed conditions on dynamic

243 response. The objective is to quantify the transverse deflection and failure mode of

244 plates as functions of loads intensity, geometric thickness and load configuration.

245 Although several input variables were considered, for brevity, this paper focused on

246 the dynamic response of 0.5mm aluminum plates under air-backed loading condition

247 in the following sections.

248 It should be noted that mechanical coupling between the tube and the water

249 column reduces the speed at which pressure pulses propagate in the fluid, cw [35]. In

250 order to quantify these effects, the pressures captured by different transducers at the

251 same tests were used to calculate the speed of sound in the water tube. Measurements

252 provided cw=1106m/s, significantly lower than the speed of sound in open water

253 1500m/s. As the decay time is controlled by the mass and velocity of projectiles,

254 5mm-thick and 12mm-thick projectiles are used to generate two different decay times:

11

Page 11 of 32
255 35us and 58us according to eq.(2). Similar to Fleck and Deshpande [21], we assume

256 that the entire impulse I0 of the shock is transmitted to the front face of the solid plate

257 to simplify the experimental analysis. In the following analysis, normalized impulses

258 according to eq.(6) are used to quantify the incident impulses. In the eq.(6), the

259 measured speed of sound in water 1106m/s and different decay times obtained by

260 different masses of strikers are used to calculate the normalized impulses. As shown

261 in Fig. 4, different pressure histories of impulsive loads are captured by pressure

262 transducers. Fig. 4 shows the pressure histories with two different decay times at

263 different initial velocity of projectiles. The predicted pressure by eq.(1) histories are in

264 good agreement with the histories in the experiments carried out with two distinct

265 decay times.

266

267 3.1 Dynamic deformation and failure modes

268 During the experimental test, two high-speed cameras were used to capture the

269 dynamic deformation of the panels. A sequence of dynamic deformations of a 0.5mm

270 plate subjected to normalized impulse 0.16 is shown in Fig. 5. In order to observe the

271 change more clearly, the amplitude of deflections for all points is amplified 2 times.

272 At the beginning of the fluid-structure interaction, circumferential dynamic plastic

273 hinges emanate from the supports to the mid-span of the plate, showing the shrinking

274 of the central red area at the plots. The midpoint reaches the highest value 8.06mm at

275 0.75ms and falls to 7.71mm after the elastic recovery, as shown in Fig. 6 (a) as well.

276 The sectional deflection profile versus time of the 0.5mm plate is plotted in Fig. 6 (a).

277 The symmetric profiles of deflections caused by the flow of dynamic hinges exhibit

278 the identical process as shown in Fig. 5. It is obvious that the postmortem
12

Page 12 of 32
279 cross-section deformation of the plate shows great agreement with the ultimate profile

280 recorded by DIC.

281 Fig. 6 (b) shows the transverse deflection history of central points on the 1.0mm

282 plates subjected to four different normalized impulsive intensities, 0.11, 0.24 and 0.25.

283 After the initially linear increase, the mid-point normal deflection remains roughly

284 constant at peak value. The histories of central deflection indicate that the effect of

285 loading rate not only affects the peak deflection, but also the response rate of

286 deflection. The response rate, as well as the maximum normal displacements,

287 increases monotonously with the increasing impulses. The panels exhibit similar rates

288 of response at the initial phase, but the responses of deflections diverge at 0.03ms and

289 0.09ms successively. This suggests that the duration of similar deflection rate

290 increases with the increase of impulsive loading. Additionally, slight elastic recoveries

291 of the deformation shown in the Fig. 5 and Fig. 6 (a) can be also observed in every

292 test after the peak value.

293 After strengthening the impulsive intensity, the failure modes of thin metallic

294 plates change from inelastic deformation to tearing and shearing, with crack initiation

295 at the mid-span or periphery of loading area. Three failure modes for metallic plates

296 have been identified by previous studies, which are divided into mode I for plastic

297 deformation with weak impulse, mode II for tearing at support with medium impact

298 intensity, and mode III for shearing failure at high strength impulsive loads. The

299 dynamic responses of inelastic deformation (mode I) are shown in Fig. 5 and Fig. 6,

300 as well as the quantified relation between dynamic deflection and impulsive

301 intensities. Rather than appearing at the support, the initial crack occurs at the

302 midpoint of the plate when the impulse exceeds the limited intensity for mode I.

303 High-speed photographic sequences of the dynamic deformation and crack initiation of
13

Page 13 of 32
304 a 0.5mm solid plate subjected to underwater impulse are shown in Fig. 7. Before the

305 initiation of central tearing, the flow of dynamic plastic hinges within the loading area

306 propagate toward the central region as shown in Fig. 5. When the stretch strain at the

307 mid-span exceeds the fracture strain at t=300µs, the tearing commences at the central

308 point of the plate. Then, the plate fails ultimately by the formation of four petals (as

309 shown in Fig. 8 as mode IIc). With the increasing impact velocity of the projectile,

310 tearing failure initiates at the periphery of the loading area. Within the range of

311 impulsive intensities, different failure modes for plates are presented in Fig. 8, mode I,

312 mode II, and the central tearing failure mode IIc. Fig. 9 shows the in-plane strain

313 histories of the midpoint and 1/4 radius point (near midpoint) at airside of the plates

314 under different tested conditions. The positive strain histories indicate that the failure of

315 the back side face sheets of structures is governed by the axial tension, regardless of the

316 differences of loading conditions.

317 3.2 Effect of thickness

318 In mode I, the transverse deflection of solid panels subjected to underwater

319 impulsive loads is sensitive to the intensity of impulses. The normalized transverse

320 deflections (  ) of plates with different thicknesses subjected to underwater impulses

321 are shown in Fig. 10 (a). In this case, the thicknesses of pistons and projectiles are

322 12mm and 5mm respectively, which means the decay times are 35µs according to

323 eq.(2). Table 2 lists the results of experiments conducted to probe the effect of

324 thickness. In the table, h ,  and I are the non-dimensional parameters obtained from

325 eq.(6).

326 For the identical thick panels, the relationship between the normal deflections

327 and impulsive intensities of the plates exhibits a good linear relation. The increasing
14

Page 14 of 32
328 thickness strengthens the impulse resistance of the plates. Three fitting curves in Fig.

329 10 (a) show similar slopes, which indicates that the thickness affects the magnitudes

330 of plastic deformation, but it nearly has no influences on the sensitivity for impulsive

331 intensities. To obtain the effect of thickness on the deflection rate, a comparison of

332 mid-point deflection history between 1.0mm-thick and 1.5mm-thick panels subjected

333 to two different impulses is shown in Fig. 10 (b). The results show that the deflection

334 rate increases with the increasing applied impulses and the thickness of panels, which

335 indicates that the thicker panel experiences longer duration of dynamic response than

336 the thinner panels at the identical applied impulses. When the normalized impulse is

337 0.21, the deflection of the 1.5mm-thick panel moves much more rapidly than the

338 1.0mm-thick panel in the direction away from the impulsive load during the rising

339 time, and all the deflections almost keep parallel eventually. As the loading intensity

340 decreases, the rate of deflection also decreases monotonically. When 1.0mm target

341 and 1.5mm target are subjected to normalized impulses 0.11 and 0.12 respectively, the

342 initial slope of the 1.5mm target is still steeper than that of the thinner plate, even

343 though the thinner one has a much larger deflection after 0.45ms. The result of the

344 comparison shown in Fig. 10 (b) indicates that the effect of thickness promotes the

345 blast resistance of monolithic aluminum plates for it not only reduces the permanent

346 plastic deformation, but also shortens the time of fluid-structure interaction by

347 increasing the response rate rapidly.

348 The structure-load-performance relation is used to achieve the quantified relation

349 between performance parameter Z (δ/R), influencing factors x (r, , h ), and load

350 properties y ( I ). A, m, n are constants specific to each load condition. The form is

351 given by

15

Page 15 of 32
(8
m n
Z  Ax y
)

352 Fig. 11(a) shows the load-structure-performance map of normalized deflections

353 (δ/L)ab of midpoints for all tests under air-backed conditions as a function of

354 normalized impulse ( I ) and panel thickness (h/R). The transverse deflections

355 increase as the applied impulses increases and decrease as the panel thickness

356 increases. For all impulses, the 1.5mm-thick panel exhibits better deflection resistance

357 than the other panels when subjected to the identical applied impulses. Fitting the

358 experimental data shown in Fig. 11 (b), the relationship between deflection in

359 air-backed plates (δ/R)ab, and normalized incident impulse ( I ) and plate thickness

360 (h/R) can be quantified by

(9
 0 .4 2
(  / R ) a b  0 .2  ( h )
0 .8 7
(I )
)

361 3.2.1 Effect of fluid-structure interaction parameter 

362 Table 3 lists the results of experiments carried out to investigate the effect of

363 fluid-structure interaction parameter on the deflection resistance of solid panels. As

364 discussed in section 2.1, the two distinct fluid-structure interaction parameters

365 controlled by the mass of strikers are 5.3 and 8.7 and the resulting decay times

366 correspond to 35µs and 58µs respectively. The results of the 0.5mm aluminum plates

367 subjected to the underwater impulsive loads with two distinct  are shown in Fig.

368 12. The central deflection exhibits a linear increase as the loading intensity increases

369 from 0.18 to 0.35, which shows a similar slope with the plates subjected to impulses

370 with FSI parameters 5.3. Compared with the deflection for  =5.3, the deflection for

371  =8.7 is ∼ 41% lower at all range of load magnitudes. When  is equal to 8.7,

16

Page 16 of 32
372 petalling failure commences when the plate is subjected to normalized impulse 0.37.

373 The quantified relation caused by Taylor parameter is expressed with the impact

374 intensities as the following form

( 10
 0 .8 4
(  / R ) a b  6 .0 5  
1 .0 1
(I )
)

375 3.2.2 Effect of loading patch size

376 The response of clamped monolithic beams to impulsive loading over a central

377 loading patch was first studied by Martin et al. [36]. They concluded that the response

378 of solid beam can be separated into three sequential phases. Two distinct patch sizes,

379 r<0.5 and r>0.5, were studied by Qiu and Deshpande [1], and the analytical model

380 was developed for the response of clamped monolithic plate and sandwich beams

381 subjected to impulses over those two kinds of central loading patches. In this paper,

382 the ratio r>0.5 was investigated experimentally by changing the inner diameter of the

383 annular clamped metallic ring. Loading over the entire span and central span, r=1 and

384 r=0.7, are compared to investigate the effect of the loading patch size, as shown in Fig.

385 1 (e). Experimental results involving in the effect of patch size are listed in Table 4.

386 Fig. 13 (a) shows transverse deflection history of midpoints when the plate is

387 subjected to the central loaded underwater impulses, where r=0.7 and  =8.7. After a

388 rapid increase at an initial time of 0.48ms (phase I), the normal deflection remains a

389 constant for about 0.30 ms (phase II) followed by the peak deflection (phase III).

390 Compared with the central deflection shown in Fig. 6 (b), a new response phase

391 (phase II) is observed when the plates are subjected to impulses over a central loading

392 patch (r=0.7). The propagation of dynamic plastic hinges can explain the discrepancy

393 between those two configurations. Initial plastic hinges emanate from the edge of

17

Page 17 of 32
394 loading and opposing propagate to plate boundary and midpoint (phase I). When the

395 patch size r>0.5, the dynamic plastic hinges reach the plate boundary before arriving

396 at the midpoint. Meanwhile, the central segment moves at a constant velocity resulted

397 from the initial impact (phase II). The plastic wave continues to propagate until it

398 reaches the midpoint of the plate and results in the maximum normal plastic

399 deflection (phase III) followed by a slight elastic recovery (phase IV). The

400 propagation of the plastic wave can be observed more obviously in Fig. 14. Failure

401 modes shown by the central loaded plate also present the difference between the two

402 distinct patch size configurations, as shown in Fig. 13 (b). The ultimate deformation

403 of central loaded plate is a dome-shape profile with smaller curvature at the mid-span

404 but larger curvature near the plate boundary than that of the fully loaded plates (r=1).

405 Transverse deflection of the monolithic plate is sensitive to the patch size of

406 loading area. When the plate is subjected to centrally loaded impulse, the amplitude of

407 plastic deformation is much larger than that of panel subjected to fully loaded impulse

408 with identical intensity, as shown in Fig. 15. As the impulse increases, the transverse

409 deflection of the midpoint grows linearly, and tearing failure occurs as the normalized

410 impulse reaches 0.35. For the 0.5mm plates, the normal deflection is ~62% larger than

411 that of panel subjected to fully loaded impulses with identical intensity. Compared to

412 the normal deflection of 0.5mm plates with 1.0mm plates subjected to loads over

413 central patch r=0.7 and r=1, an obvious difference can be observed from Fig. 15 and

414 Fig. 10 (a). When the plate is fully loaded, the thicker plate only improves the

415 deflection resistance ~20%. However, this improvement increases ~51% when the

416 two kinds of plate are subjected to impulsive loads over central patch r=0.7. The

417 relationship between normalized transverse deflection, patch size and impulse can be

418 given by
18

Page 18 of 32
(1
 1 .9 9
(  / R ) a b  0 .9 5  r
0 .9 9
(I ) 1)

419 3.3 Effect of water-backed configuration

420 Water-backed configuration is a common condition for structures in large naval

421 structures, such as keels, rudders, propeller blades and underwater pipelines. A

422 comparison of results between air-backed and water-backed panels reveals significant

423 discrepancies in deformation and failure mechanism. Instead of transmitting the

424 impulse to the supports, the water-backed plate transmits most of the impulse into the

425 backside water by deformation. The presence of water on both sides of plates prevents

426 large scale bending and absorbs much more initial energy by the backside water

427 column. To quantify the differences under these two conditions, a comparative study

428 was carried out experimentally.

429 In the water-backed condition, another 150mm long water tube is horizontally

430 mounted at the backside, as shown in Fig. 1 (b). In this case, the 12mm-thick (0.22kg)

431 projectiles are selected, which means that the decay time θ and fluid-structure

432 interaction parameter Ψ are θ=58µs and Ψ=8.7 respectively. The results for

433 water-backed tests are shown in Table 5.

434 Fig. 16 shows the central deflection of the water-backed plate with thicknesses of

435 0.5mm and 1.0mm subjected to underwater impulsive loads. Due to the resistance of

436 backside water, the overall deflections of plates are limited significantly. Comparing

437 the central deflections under air-backed and water-backed condition, the 0.5mm-thick

438 panels undergo ~53% lower central deflections under water-backed condition than

439 that of the same thick panels under air-backed condition. As the loading intensity

440 increases from I =0.20 to I =0.45, the trends of deflection for both 0.5mm-thick and

19

Page 19 of 32
441 1.0mm-thick plates exhibit linear relationships. Under water-backed condition, the

442 discrepancy of transverse deflection between 0.5mm-thick and 1.0mm-thick panels is

443 small at lower impulsive intensity, but it increases rapidly with the increase of

444 impulses. The relationship among normalized deflection under water-backed plates,

445 incident normalized impulse, and thickness of plates is quantified by

 1 .2 5 ( 12 )
(  / L ) w b  0 .0 2  ( h )
2 .6 4
(I )
446 The presence of water not only restricts the transverse deflection but also affects

447 the failure mechanism of the clamped plates. Fig. 17 exhibits a cross-section profile of

448 the 0.5mm panel subjected to normalized impulse 0.28 under water-backed condition,

449 which differs from the sectional deflection profiles under air-backed condition

450 obviously in the shape and amplitude. There is an obviously corrugated-shape

451 circumferential bulge in the direction away from the impulsive loading at the plate

452 support area and a dented deformation at the circumferential boundary of loads. The

453 central area presents a conical deformation which is different from all of the deform

454 profiles presented in the air backed tests. The slightly corrugated-shape deformation is

455 observed at all postmortem images obtained from water-backed tests. Once the panel

456 is subjected to impulse, the incident energy is converted into the deformation energy

457 in the plate and then transmits most of the energy into backside water in a short

458 amount of time. The plate compresses the backside water by transverse deformation at

459 the initial time. Due to the incompressibility of the water, the compressive wave

460 transmitted from the plate impinges on the plate again after it is reflected from the

461 boundary of the backside water tube, to make an overall reverse deflection on the

462 plates. The process of deformation and reverse deformation of plates lead to the

463 corrugated-shape deformation near the loading boundary.

20

Page 20 of 32
464 4. Concluding remarks

465 Dynamic response and failure of 5A06 aluminum alloy solid plates subjected to

466 water-based impulsive loadings have been evaluated experimentally in this paper. The

467 effect of panel thickness, fluid-structure interaction parameter, loading patch size and

468 intensity of underwater impulse on the dynamic deformation, failure modes and

469 associated mechanisms of the solid plate in air-backed and water-backed conditions

470 are examined respectively.

471 This study has yielded experimental data on the dynamic response of blast

472 resistance of monolithic metallic plates subjected to water-based impulses. The

473 dynamic deformation of solid plates indicates that thickness of panels contributes

474 limited influences on the response of plates. The response rate increases with the

475 increase of applied impulses and thickness of panels. In the central loading condition

476 (patch size r=0.7), an additional response phase is exhibited by the panels due to the

477 different propagations of dynamic plastic hinges, which is the primary factor on distinct

478 profile of deformation. The failure modes of solid panels includes inelastic deformation,

479 tearing at central region and tearing at supports with the increasing impulsive loads.

480 The blast-resistant performance as a form of inelastic deflection is quantified

481 considering the effect of plate thickness, Taylor’s FSI parameter, and patch size of

482 loading under the air-backed condition. The parallel linear relations for different

483 thicknesses indicate that the transverse deflections increase as the applied impulses

484 increases and decrease as the panel thickness increases. The thickness of panel

485 influences only the amplitudes of deflection rather than the sensitivity to impulsive

486 intensities. The loading patch size involved in this paper is verified as the most

487 sensitive effect on the deflection resistance of panels. In the central loading condition,

21

Page 21 of 32
488 the peak deflection increases ~62% for 0.5mm panels, and discrepancy between

489 0.5mm-thick and 1.0mm-thick panels has been amplified from 20% to 51%. The

490 deflection resistance of panels undergoes obvious improvement when the

491 fluid-structure interaction parameter increases from 5.3 to 8.7, e.g. the deflections of

492 0.5mm panels decrease ~41% at bigger fluid-structure interaction parameter.

493 Due to the bending resistance of backside water, solid panels experience much

494 lower transverse deflection under water-backed condition than the panels under

495 air-backed condition, e.g. the deflections of 0.5mm panels decrease ~53% under

496 water-backed condition. The discrepancy of deflections between 0.5mm and 1.0mm

497 shows an increasing trend within the range of impulses rather than the almost parallel

498 relation under air-backed conditions. The cross-section profile of the plate subjected

499 to water-backed condition shows a circumferentially corrugated-shape deformation at

500 the periphery of loads and a conical deformation at the mid-point, which are different

501 from the deformation shown in air-backed condition.

502 The quantitative structure-loads-performance relationship in terms of transverse

503 deflection as functions of impulsive loads, geometric thickness, fluid-structure

504 parameter and patch size of loading, are investigated respectively, as shown in Table 6.

505 The insight obtained in this paper provides guidelines for the design of metallic

506 structures for which response to water-based impulsive loading is an important

507 consideration. Finally, it should be noted that the relationships described in this paper

508 are applicable considering the structural attributes and loading conditions.

509 Acknowledgments

510 The authors would like to thank the National Natural Science Foundation of China

511 (No.: 11372088, 51509115) for supporting the present work.


22

Page 22 of 32
512 References

513 [1] X. Qiu, V.S. Deshpande, N.A. Fleck, Impulsive loading of clamped monolithic and sandwich
514 beams over a central patch, Journal of the Mechanics and Physics of solids, 53 (2005)
515 1015–1046.
516 [2] V.S. Deshpande, A. Heaver, N.A. Fleck, An underwater shock simulator, Proc. R. Soc. Lon.,
517 Ser-A 462 (2006) 1021–1041.
518 [3] Z. Xue, J.W. Hutchinson, Preliminary assessment of sandwich plates subject to blast loads, Int. J.
519 Mech. Sci. , 45 (2003) 687–705
520 [4] S. Avachat, M. Zhou, Response of submerged metallic sandwich structures to underwater
521 impulsive loads, Journal of Mechanics of Materials and Structures, 10 (2015 ) 17–41.
522 [5] A. Schiffer, V.L. Tagarielli, The response of rigid plates to blast in deep water: fluid–structure
523 interaction experiments, Proceedings of the royal sociaty: A, (2014).
524 [6] A. Schiffer, V.L. Tagarielli, The one-dimensional response of a water-filled double hull to
525 underwater blast: experiments and simulations, International Journal of Impact Engineering,
526 63 (2014) 177–187.
527 [7] A. Schiffer, V.L. Tagarielli, The dynamic response of composite plates to underwater blast:
528 theoretical and numerical modelling, International Journal of Impact Engineering, 70 (2014)
529 1–13.
530 [8] K.P. Dharmasena, H.N.G. Wadley, K. Williams, Z. Xue, J.W. Hutchinson, Response of metallic
531 pyramidal lattice core sandwich panels to high intensity impulsive loading in air, Int. J. Impact
532 Eng., 38 (2011) 275–289.
533 [9] K.P. Dharmasena, D.T. Queheillalt, H.N.G. Wadley, P. Dudt, Y. Chen, D. Knight, A.G. Evans,
534 V.S. Deshpande, Dynamic compression of metallic sandwich structures during planar
535 impulsive loading in water, Eur.J. Mech. A Solids, 29 (2010) 56–67.
536 [10] S. Avachat, M. Zhou, High-speed digital imaging and computational modeling of dynamic
537 failure in composite structures subjected to underwater impulsive loads, International Journal
538 of Impact Engineering, 77 (2015) 147-165.
539 [11] Y.P. Zhao, Suggestion of a new dimensionless number for dynamic plastic response of beams
540 and plates, Archive of Applied Mechanics, 68 (1998) 524-538.
541 [12] N. Jones, Structural Impact, Cambridge University Press,UK, Edition 2, ISBN:
542 9781107010963,, (2012) 1-604.
543 [13] N. G.N., M. J.B, Deformation of thin plates subjected to impulsive loading - a review. Part II:
544 Experimental Studies, Int J Impact Eng,, 8 (1989) 171-186.
545 [14] G.N.Nurick, J.B.Martin, Deformation of thin plates subjected to impulsive loading - a review.
546 Part I: Theoretical considerations, Int J Impact Eng,, 8 (1989) 159-169.
547 [15] H.D. Espinosa, S. Lee, N. Moldovan, A novel fluid structure interaction experiment to
548 investigate deformation of structural elements subjected to impulsive loading, Exp. Mech., 46
549 (2006) 805–824.
550 [16] G.J. McShane, V.S. Deshpande, N.A. Fleck, The Underwater Blast Resistance of Metallic
551 Sandwich Beams With Prismatic Lattice Cores, J. Appl. Mech. (ASME), 74 (2006) 352–364.
552 [17] A. Schiffer, V.L. Tagarielli, One-dimensional response of sandwich plates to underwater blast:
553 Fluid-structure interaction experiments and simulations, International Journal of Impact
554 Engineering 71 (2014) 34-49
555 [18] V.S. Deshpande, N.A. Fleck, Isotropic constitutive models for metallic foams, Journal of the
556 Mechanics and Physics of Solids, 48 (2000) 1253-1283.
557 [19] V.S. Deshpande, N.A. Fleck, One-dimensional response of sandwich plates to underwater
558 shock loading, J. Mech. Phys. Solids, 53 (2005) 2347–2383.
559 [20] Z. Xue, J.W. Hutchinson, Constitutive model for quasi-static deformation of metallic sandwich
560 cores, Int. J. Numer. Methods Eng. , 61 (2004) 2205–2238.
561 [21] N.A. Fleck, and Deshpande, V. S., The Resistance of Clamped Sandwich Beams to Shock
562 Loading, ASME J. Appl. Mech., 71 (2004) 386–401.
563 [22] N. Jones, Structural impact, Cambridge university press, 2011.
564 [23] G.N. Nurick, G.C. Shave, The deformation and tearing of thin square plates subjected to
565 impulsive loads- an experimental study, Int. J. Mech. Sci. , 18 (2000) 99–116.
23

Page 23 of 32
566 [24] Y.W. Lee, T. Wierzbicki, Fracture prediction of thin plates under localized impulsive loading.
567 part II: discing and petalling, Int. J. Impact Eng., 31 (2005b) 1277–1308.
568 [25] Y.W. Lee, T. Wierzbick, Fracture prediction of thin plates under localized impulsive
569 loading.part I: dishing, Int. J. Impact Eng. , 31 ( 2005a) 1253–1276.
570 [26] V.H. Balden, G.N. Nurick, Numerical simulation of the post-failure motion of steel plates
571 subjected to blast loading, Int. J. Impact Eng., 32 (2005) 14–34.
572 [27] S. Kazemahvazi, D. Radford, V.S. Deshpande, N.A. Fleck, Dynamic failure of clamped
573 circular plates subjected to an underwater shock Journal of the Mechanics of Materials and
574 Structures, 2 (2007) 2007–2023.
575 [28] J. Zamani, K.H. Safari, A.K. Ghamsari, A. Zamiri, Experimental analysis of clamped AA5010
576 and steel plates subjected to blast loading and underwater explosion, J. Strain Analysis, 46
577 (2010) 201-212.
578 [29] W. Huang, W. Zhang, P. Ren, Z.T. Guo, N. Ye, D.C. Li, Y.B. Gao, An Experimental
579 Investigation of Water-Filled Tank Subjected to Horizontal High Speed Impact, Experimental
580 Mechanics, 55 (2015) 1123–1138.
581 [30] Z. Xue, J.W. Hutchinson, A Comparative Study of ImpulseResistant Metallic Sandwich Plates,
582 Int. J. Impact Eng., 30 (2004) 1283–1305.
583 [31] Y. Liang, A.V. Spuskanyuk, S.E. Flores, D.R. Hayhurst, J.W. Hutchinson, R.M. McMeeking,
584 A.G. Evans, The response of metallic sandwich panels to water blast, J Appl Mech, 74 (2007)
585 81–99.
586 [32] G.I. Taylor, Scientific papers, III: Aerodynamics and the mechanics of projectiles and
587 explosions, edited by G. K., Batchelor, Cambridge University Press, (1963).
588 [33] A. Schiffer, V.L. Tagarielli, The response of circular composite plates to underwater blast:
589 Experiments and modelling, Journal of Fluids and StructuresVolume 52 ( 2015) 130–144.
590 [34] G.R. Johnson, W.H. Cook, Fracture characteristics of three metals subjected to various strains,
591 strain rates, temperatures and pressures, Eng. Fract. Mech., 21 (1985) 31–48.
592 [35] D.J. Korteweg, Uber die Fortpflanzungsgeschwindigkeit des Schalles in elastischen Rohren,
593 Annals of Physics 5(1878.) 525–542.
594 [36] J.B.Martin, P.S. Symmonds, Mode approximations for impulsively-loaded rigid–plastic
595 structures, J. Eng. Mech.: ASCE EM5, (1966) 43–66.
596
597
Pressure Pressure Pressure
transducer Simulator tube transducer transducer Piston
(a) Camera Specimen Projectile Gas
gun
Light 66mm
source
Water
Camera
500mm
Pressure Pressure Pressure
transducer Simulator tube transducer transducer Piston
Projectile
(b)
Water Water

150mm 500mm

Rubber Ring Loading area

Annular Ring
(c)
33mm 48mm 48mm
33mm
Impulse 80mm 80mm

Clamped
Bolt Edge clamped region

598 (d) (e)

24

Page 24 of 32
599 Fig. 1.Sketch of the apparatus employed in this study to explore the response of (a) air-backed
600 condition, (b) water-backed condition, (c) side view of clamped panel: (d) specimen in air-backed
601 condition (e) specimen in central patch r=0.7 loading condition.
602

603 (a) (b)

604 Fig. 2. Geometry of the smooth specimen and post-test photographs for uniaxial tensile test (a)
605 room and high temperature quasi-static tensile specimens and (b) dynamic tensile specimens
606
400 400
(a) (b) o
20 C
o
100 C
Engineering stress (MPa)
Engineering Stress (MPa)

o
300 300 150 C
o
200 C
o o
0 250 C
200 o
200
0
o
45
o
100 45 100
o
90
o
90
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.16 0.32 0.48 0.64

607 Engineering Strain Engineering strain

500 (c)
Engineering stress (MPa)

400

300

-3
200 1.2510 /s
713.72/s
1493.85/s
100 1733.31/s
1968.35/s
fitting curve
0
0.00 0.03 0.06 0.09 0.12 0.15

Engineerig Strain
608

609 Fig. 3. Engineering stress-strain curves of 5A06 aluminum(a) quasi-static tensile,(b)high


610 temperature tensile, and (c) high strain rate tensile.
611

25

Page 25 of 32
120 250
Reservior pressure 0.6 MPa (a) Reservior pressure 2.0 MPa (b)
Projectile velocity 75.8m/s Projectile velocity 147.5m/s
200
Pressure (MPa) 90 I=0.13 s I=0.24 s

Pressure (MPa)
150

60 Experiment
Experiment 100 Prediction
Prediction
30 50

0
0
-50

0 100 200 300 400 500


0 100 200 300 400 500
Time (s)
Time (s)
612

300
90
Reservior pressure 0.5 MPa (c) Reservior pressure 3.0 MPa
(d)
Projectile velocity 64.6m/s 250
Projectile velocity 162.6m/s
I=0.17 s I=0.44 s

Pressure (MPa)
200
Pressure (MPa)

60

150

30 Experiment 100 Experiment


Prediction Prediction
50

0
0

-50
0 100 200 300 400 500 0 100 200 300 400 500
Time (s) Time (s)
613
614 Fig. 4. Experimentally measured and predicted pressure histories for four different projectile
615 velocities,(a)(b) decay time was 35µs ,(c)(d) decay time was 58µs.
616

0.00ms 0.09ms 0.24ms

0.33ms 0.45ms 0.75ms

617
618 Fig. 5. Evolution of normal deformation for 0.5mm monolithic aluminum plate subjected to
619 normalized impulse 0.16.
620

26

Page 26 of 32
9.0 11.0
dynamic deformation 0.75ms
post-test result I=0.25
Normal deflection (mm)
7.5

Normal deflection (mm)


1.05ms post-test 8.8 I=0.24
0.45ms
6.0
6.6
4.5 0.33ms
I=0.11
4.4
3.0 0.24ms

1.5 2.2
0.09ms
0.0 (a) (b)
0.0
0 10 20 30 40 50 60 0.0 0.3 0.6 0.9 1.2
Cross-section profile (mm) Time (ms)
621

622 Fig. 6. (a) Deflection profile evolution of and post-tested profile for 0.5mm plate, I =0.16, θ=35µs,
623 and (b) normal deformation history of midpoints, 1.0mm plates.
624

0 us 120us 180us

210us 300us 420us


625
626 Fig. 7. Sequence of high-speed photographs showing the central tearing failure in 0.5mm
627 monolithic plate subjected to underwater impulsive loading with I =0.37, θ=58us.

Mode I Mode IIc Mode II

628
27

Page 27 of 32
629 Fig. 8. Typical failure modes of solid metallic plates.

4.5 6
4.0
(a) (b)
5
3.5
3.0 4
( %)

2.5

( %)
3
2.0
1.5 2
1.0
1
0.5
0.0 0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
Time(ms) Time(ms)
630
631 Fig. 9 Strain histories at different points on the back side of plates, (a) h=0.5mm, normalized
632 impulse 0.16 with patch size r=1.0, and (b) h=1.0mm, normalized impulse 0.39 with patch size
633 r=0.7.
634
0.35
9.0
(a) I=0.21
(b)
Normalized deflection (L)

0.30
7.5
Transverse deflection 

I=0.21
0.25
6.0
I=0.11
0.20
4.5
0.15 I=0.12
3.0
0.5mm
0.10 1.0mm
1.5mm 1.5 1.0mm
fitting curve
0.05
1.5mm
0.0
0.05 0.10 0.15 0.20 0.25 0.30 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Normalized impulsive (I) Time (ms)
635

636 Fig. 10 (a)Transverse deflections versus applied impulses, and (b) a comparison of mid-point
637 deflection rate.

638

639 Fig. 11 Normalized deflection as a function of normalized impulses and panel thickness, (a) map of
640 load-structure-performance, and (b) 3D bar picture of experimental data and fitting surface.
641

28

Page 28 of 32
0.4
crack
Normalized deflection (L)
0.3

0.2
s,
s,
(a) fitting curve
0.1
0.1 0.2 0.3 0.4
Normalized Impulse (I)
642
643 Fig. 12 Normalized deflection as a function of normalized impulses and fluid-structure interaction
644 parameter, (a) linear relations of load-structure-performance, and (b) 3D bar picture of
645 experimental data and fitting surface.
646
h=1.0mm, I =0.39

20
18
phase I
Transverse deflection 

16
14
12 phase IV
phase II phase III
10
Central loaded plate, r=0.7
8
1.0mm,I=0.45
6 1.0mm,I=0.39
4
2
(a)
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 Entire loaded plate, r=1
Time (ms) (b)
647

29

Page 29 of 32
648 Fig. 13. (a)mid-points transvers deflection history of the central loaded plates, and (b) a comparison
649 of deflection profile between r=0.7 and r=1 loading conditions.

0.00ms 0.15ms 0.50ms

0.70ms 1.15ms 1.50ms

650
651 Fig. 14. Evolution of normal deformation for 1.0mm central loaded plate subjected to normalized
652 impulse 0.39.
653

0.7
0.5mm,r=1.0
0.5mm,r=0.7 crack,I=0.35
Normalized deflection (R)

0.6
1.0mm,r=0.7
0.5 fitting curve

0.4

crack,I=0.37
0.3

0.2

0.1 (a)
0.00 0.10 0.20 0.30 0.40
Normalized impulse (I)
654
655 Fig. 15 Normalized deflection as a function of normalized impulses and loading patch size r, (a)
656 linear relations of load-structure-performance, and (b) 3D bar picture of experimental data and
657 fitting surface.
658

0.45
Normalized deflection (L)

0.5mm,air-backed
0.5mm,water-backed
1.0mm,water-backed
0.36 fitting curve -
crack,I=0.20

0.27

0.18

0.09

(a)
0.00
0.18 0.28 0.37 0.46
Normalized Impulsive(I)
659

30

Page 30 of 32
660 Fig. 16. Normalized deflection as a function of normalized impulses and normalozed thickness
661 under water-backed condition, (a) a comparison of linear relations of load-structure-performance in
662 two different loading conditions, and (b) 3D bar picture of experimental data and fitting surface.
663
15
water-backed

Transverse deflection (mm)


air-backed
12
I=0.28

0
Front side Back side -50 -40 -30 -20 -10 0 10 20 30 40 50
(a) (b) Cross-section profile (mm)
664

665 Fig. 17. The postmertem images of 0.5mm panel subjected to I =0.28 under water-backed
666 condition, (a) failure modes of the panel, and (b) a comparison of post-test deflection of 0.5mm
667 panels subjected to identical impulse under two loading conditions.
668
669 Table 1. Mechanical properties of 5A06 aluminum alloy
Material Young’s Density A B C n m
modulus (kg/m3) (MPa) (MPa)
(GPa)
5A06 74 2780 167.0 458.7 0.44 0.02 2.3
670
671 Table 2 Results of experiments for effect of thickness
Contact Normalized Decay Initial Normalized Normalized Failure
condition thickness time velocity impulse deflection mode
h (µs) (m/s) I 
air-acked,r=1.0 0.015 35 68.4 0.11 0.19 I
35 96.8 0.16 0.23 I
35 137.6 0.22 0.31 I
35 101.6 0.16 0.23 I
air-acked,r=1.0 0.030 35 70.8 0.11 0.15 I
35 158.5 0.25 0.29 I
35 129.3 0.21 0.23 I
35 147.9 0.24 0.24 I
air-acked,r=1.0 0.045 35 58.7 0.07 0.05 I
35 136.8 0.22 0.17 I
35 128.9 0.21 0.21 I
35 76.2 0.12 0.10 I
672

673
674 Table 3 Results of experiments for effect of thickness fluid-structure interaction parameter
Contact Normalized Decay Initial Normalized Normalized Failure
condition thickness time velocity impulse deflection mode
31

Page 31 of 32
h (µs) (m/s) I 
air-bcked,r=1.0 0.015 58 140.5 0.37 failed IIc
58 102.0 0.27 0.29 I
58 132.6 0.36 0.35 I
58 76.8 0.19 0.14 I
675
676 Table 4 Results of experiments for effect of patch size
Contact condition Normalized Decay Initial Normalized Normalized Failure
thickness time velocity impulse deflection mode
h (µs) (m/s) I 
air-acked,r=0.7 0.015 58 75.2 0.20 0.41 I
58 30.9 0.04 0.10 I
58 107.8 0.29 0.57 I
58 127.5 0.35 failed II
air-acked,r=0.7 0.030 58 104.4 0.28 0.39 I
58 166.9 0.45 0.49 I
58 145.5 0.39 0.45 I
58 119.0 0.32 0.36 I
58 76.2 0.21 0.14 I
677
678 Table 5 Results of experiments for effect of water-backed configuration
Contact Normalized Decay Initial Normalized Normalized Failure
condition thickness time velocity impulse deflection mode
h (µs) (m/s) I 
water-backed 0.015 58 104.6 0.28 0.16 I
58 116.6 0.31 0.20 I
58 77.2 0.20 0.04 I
58 105.8 0.28 0.15 I
0.030 58 102.0 0.27 0.09 I
58 146.5 0.39 0.15 I
58 168.1 0.44 0.21 I
58 80.5 0.21 0.03 I
679
680 Table 6 Summary of material-structure-property relationships for clamped aluminum plates.
Figure Contact condition Effect factor Structure-performance relations
Fig. 11 air-backed h (  / R ) a b  0 .2  ( h )
 0 .4 2
(I )
0 .8 7

Fig. 12(b) Ψ (  / R ) a b  6 .0 5  
 0 .8 4
(I )
1 .0 1

Fig. 15(b) r  1 .9 9
(  / R ) a b  0 .9 5  r
0 .9 9
(I )
Fig. 16(b) water-backed  1 .2 5
(  / L ) w b  0 .0 2  ( h )
2 .6 4
h (I )

681
682

32

Page 32 of 32

You might also like