A statistical analysis of nanocavities replication applied to injection moulding

A statistical analysis of nanocavities replication applied to injection moulding

ARTICLE IN PRESS ICHMT-03537; No of Pages 10 International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 001 067 Contents lists availa...

3MB Sizes 0 Downloads 53 Views

ARTICLE IN PRESS ICHMT-03537; No of Pages 10 International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 001

067

Contents lists available at ScienceDirect

002 003

069

004

International Communications in Heat and Mass Transfer

005 006

070 071 072

007

journal homepage: www.elsevier.com/locate/ichmt

073

008

074

009

075

010

076

011

013 014

077

A statistical analysis of nanocavities replication applied to injection moulding

PR OO F

Q1 012

015 016

Q2 017 018

Q3 019 020 021

J. Pina-Estany a, * , C. Colominas b , J. Fraxedas c , J. Llobet d , F. Perez-Murano d , J.M. Puigoriol-Forcada a , D. Ruso c,1 , A.A. Garcia-Granada a a

IQS-Universitat Ramon Llull, Via Augusta, 390, 08017 Barcelona, Spain b Flubetech SL, Carrer Montsià, 23, 08211 Castellar del Vallès, Spain c Catalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC, The Barcelona Institute of Science and Technology, Campus UAB, Bellaterra, 08193 Barcelona, Spain d Institut de Microelectrònica de Barcelona (IMB-CNM, CSIC), Campus Universitat Autònoma de Barcelona, C/Til • lers, Cerdanyola del Vallès, Barcelona, Spain

022 023 024 025

A R T I C L E

I N F O

A B S T R A C T

026

028

031 032 033 034

TE

030

Keywords: Nanoscale simulation AFM Heat transfer Injection moulding CFD Submodeling

EC

029

035 036 037 038

RR

039 040 041 042

1. Introduction

045

047 048 049 050 051 052 053 054 055

The demand of products with nanostructured surfaces is increasing due to the wide field of applications. The applications cover the medical and pharmaceutical industry with lab-on-chips devices (Oh [1]) and surfaces with self-cleaning (Yoo [2]) and antimicrobial surfaces (Kim [3]); environment applications to purified water(Baruah [4]) and air and to remediate the soil; optical applications for achieving antireflection (Christiansen [5]), transmission and contrast enhancements, etc. Injection moulding is a manufacturing process with a high cost of entry but a low production cost (Fagade [6]). Therefore, it is a

UN

046

CO

043 044

056

059 060 061 062 063

066

080 081 082 083 084 085 086 087 088 089 090 091

093 094 095 096 097 098 099 100 101 102 103 104 105

107 108 109

promising process for obtaining large series of plastic parts with nanostructured surfaces. However, achieving a good replication is not trivial because the obtained replication index depends on many factors, among them process factors (mould and polymer temperatures, filling time, holding pressures, viscosity of the polymer, etc.) and geometrical factors (depth, wide, aspect ratio of the cavity, etc.). The purpose of this paper is to give light on how -and how muchthese factors affect the replication and, hence, avoid failures like the one showed in the Fig. 1, where we show how a nanocavity was not successfully replicated by the polymer. The study is done both from a theoretical point of view using computer fluid dynamics but also from an experimental point of view.

110 111 112 113 114 115 116 117 118 119 120 121 122 123

Abbreviations: AFM , Atomic Force Microscope; CFD , Computational Fluid Dynamics; FIB , Focused Ion Beam; SEM , Scanning Electron Microscopy; VOF , volume of fluid. * Corresponding author. E-mail address: [email protected] (J. Pina-Estany). 1

Present address: Flubetech SL, Carrer Montsià, 23, 08211 Castellar del Vallès, Spain.

2. Material and methods This section is divided in two subsections, the first one referring to the simulation setup (Section 2.1) and the second one related to the experimental setup (Section 2.2).

064 065

079

106

057 058

078

092

The purpose of this paper is to investigate both theoretically and experimentally how nanocavities are replicated in the injection moulding manufacturing process. The objective is to obtain a methodology for efficiently replicate nanocavities. From the theoretical point of view, simulations are carried out using a submodeling approach combining Solidworks Plastics for a first macrosimulation and Fluent solver for a subsequent nanosimulation. The effect of the four main factors (melt temperature, mould temperature, filling time and cavity geometry) are quantified using an statistical 24 factorial experiment. It is found that the main effects are the cavity length, the mould temperature and the polymer temperature, with standardized effects of 5, 3 and 2.6, respectively. Filling time has a negative 1.3 standardized effect. From the experimental point of view, Focused Ion Beam technique is used for mechanizing nanocavities in a steel mould. The replication achieved in polycarbonate injection is quantified using an Atomic Force Microscope. It is observed how both the geometry and the position of the cavities in the mould affect its replication. © 2016 Elsevier Ltd. All rights reserved.

Available online xxxx

D

027

Q4

068

124 125 126 127 128 129 130

http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003 0735-1933/© 2016 Elsevier Ltd. All rights reserved.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

131 132

ARTICLE IN PRESS 2

J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 199

134

200

135

201

136

202

137

203

138

204

139

205

140

206

141

207

142

208

143

209

144

210

145

211

PR OO F

133

146

Fig. 3. Macro simulation, the graduates tones in blue indicate the polymer advance. It is considered that the fluid goes through the cavity without noticing its presence.

148 149 150 151 152 153 154 155 156

Fig. 1. AFM measurement of a failed attempt to fill a nanocavity. Even though the cavity is 900nm depth, the polymer penetrates only 50nm. The reason is that, due to the big area/volume ratio of the nanoworld, the polymer reaches its no flow temperature far before reaching the end of the cavity. This paper aims to give light into how the process and geometrical parameters should be adjusted in order to improve the replication.

157 158

164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187

TE

163

EC

162

When speaking about simulation of fluids dynamics at the nanoscale, the first question that arises is whether the continuum fluid approach is valid for such scale or not. The small cavities we simulate and experimentally fill with polymer are of around 2 lm • 10lm • 200nm = 4 • 10 −18 m3 , using the polycarbonate density (1200 kg/m3 ) and its molar mass (254.3 g/mol) we calculate that each cavity is filled with around 4 • 1011 polycarbonate units. Because of this elevated number of polycarbonate units in each cell and because the filled cavities are larger than 10 molecular diameters, which is, according to the literature (Eijkel [7]), the threshold where deviations from the classical continuum theory are observed due to the quantisation of molecular layers, the continuum approach is considered valid for this application. In consequence, the following calculations are carried using the conventional Computational Fluid Dynamics approach and the Finite Volume Method for discretizing the Navier-Stokes equations.

RR

161

2.1. Simulation setup

CO

160

1. Macro simulation first stage: A first conventional simulation of mould filling is carried out in the geometry shown in the Fig. 2. The simulations are done in this geometry because it is the geometry that was used for manufacturing the mould used for the Experimental results section. In the first stage it is considered that the fluid goes through the cavity without noticing its presence (see Fig. 3). The reason

UN

159

The first problem that arises when aiming to simulate how a polymer flows and fills a mould with a nanostructured pattern is the meshing problem. Meshing the whole mould with a nanoscale mesh would lead to a computationally unaffordable calculation, as a fast exercise: a part of (50 • 50 • 3)mm3 meshed with hexahedral mesh of 1 mm has up to 7500 elements and its calculation takes in a personal computer (Intel(R) Core(TM) i5-3320 M CPU @ 2.60 GHz) around 600 s. If the model is meshed with quadrilateral elements of 1 nm per vertex we have instead of 7500 elements up to 7.5 • 1030 elements, what leads to a calculation time of 1.9 • 1022 years without even taking into account the time step reduction necessary for maintaining the Courant number below a reasonable margin. And, apart from the time problem, no computer has enough memory for handling this huge amount of elements. A logical solution to this problem seems to be to refine the mesh around the cavities, but this leads to unphysical results due to the too big ratio of scales as already stated in the literature (Tofteberg, [8], Yu [9]). We worked around this problem by using a submodeling approach. The submodeling approach –multiscale simulation according to some authors– consists in carrying out two simulations consecutively. The first is a macro simulation carried out using Solidworks Plastics and the second is a nanoscale simulation carried out using CFD. In order to link the two simulations, the boundary conditions are transferred from the first to the second simulation:

D

147

212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253

188

254

189

255

190

256

191

257

192

258

193

259

194

260

195

261

196 197 198

Fig. 2. CAD model used for the macro simulation first stage and corresponding to the part manufactured using injection moulding as exposed in the Experimental setup section.

Fig. 4. Nano simulation, the graduated tones in blue indicate the polymer advancing into the nanocavity until it reaches the glass transition temperature. The boundary conditions are indicated.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

262 263 264

ARTICLE IN PRESS J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

3 331

266

332

267

333

268

334

269

335

270

336

271

337

272

338

273

339

274

340

275

341

276

342

277

343

PR OO F

265

278 279 280 281

283 284 285

289 290 291 292 293 294 295 296

is that the mark is so small that the polymer goes through it with around 0.15ms for a normal injection. From this macro simulation we extract the pressure profile and the melt front temperature of the polymer when it contacts for the first time the cavity. Furthermore, due to the small size of the nanocavity it is also considered that the cavity presence does not affect the bulk flow. 2. Nanosimulation: Once the polymer has gone through the mark, the pressure starts pushing the polymer into the mark (see Fig. 4). This stage gets to an end when the polymer solidifies due to the heat transfer with the walls.

297

299 300

The nanosimulation is stopped when the polymer reaches its glass transition temperature (Tg ) in all the region inside the cavity. Next, we expose the nanosimulation setup:

EC

298

301 302

306 307 308 309 310 311 312 313 314

RR

305

2.1.1. Models The models used are the volume of fluid model, the energy and the solidification/melting model. The VOF method tracks the interface evolution with an Eulerian approach, i.e., without modifying the mesh during the calculation. The explicit approach is used in order to activate the Geometric Reconstruction scheme and obtain hence a sharper interface. Laminar flow is considered due to the low Reynolds number. The energy model is used with the purpose to account for the heat transfer between the polymer and the walls. The solidification and melting model is used in order to model the solidification of the fluid due to the heat transfer with the walls.

CO

304

315 316 317 318

348 349 350 351

2.1.3. Materials Three materials are involved in the simulation: polymer, air and steel: • Polymer: – Density = 1200 kg/m3 . Density dependance of the fluid with the temperature and pressure is not taken into account for simplification purposes. When considering the error introduced by making this simplification, we identify in the PVT diagram that, if the density is taken as the average between the initial and the final conditions, the introduced error is less than a 5%. – Viscosity The viscosity of the polymer is modeled using the Cross-WLF model. The Cross-WLF models the dependance of the polymer with the temperature (viscosity inversely proportional to temperature) and with the shear rate (viscosity inversely proportional to shear rate). For the polycarbonate, the viscosity vs. shear rate for different temperatures is shown in the Fig. 5. The equations modeling the Cross-WLF are:

˙ = g(T, c)

1+

g0 (T)c˙ t∗

(1)

328 329 330

357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372

374 375 376

380 381 382 383 384 385

387

Q5

388

Table 1 Cross-WLF constants for PC.

389

Constant

Value

Unit

A1 A2 D1 D2 t∗ n

8.4 246.8 462 573 8437056 0.116

– K Pa s Pa s Pa –

390

325

327

356

386

322

326

355

379

321

324

354

378

320

323

353

377

2.1.2. Mesh A nearly structured quadrilateral mesh of 70000 elements is used.

319

352

373

g0 (T)

UN

303

347

D

288

346

TE

287

345

Fig. 5. Cross-WLF model for modeling viscosity as a function of temperature and shear rate for polycarbonate.

282

286

344

391 392 393

Fig. 6. Different approaches for fluid velocity at the wall. A perfect slip approach was used for the walls in order to avoid the unphysical results due to the contact line discontinuity of the Navier-Stokes equation.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

394 395 396

Q6

ARTICLE IN PRESS 4

J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx 463

398

464

399

465

400

466

401

467

402

468

403

469

404

470

405

471

406

472

407

473

408

474

409

475

PR OO F

397

410 411 412 413 414 415 416

Fig. 7. Pressure that would be generated inside the cavity due to compression of the air (blue) and 1 bar pressure considered in the simulations with pressure outlet boundary condition (green). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

417 418 419 420 421 422

g0 (T) = D1 e

−A1 (T−D2 ) A2 +(T−D2 )

(2)

429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446

TE

428

EC

427

RR

426

CO

425

Where g is the melt viscosity of the polymer, g0 is the zero shear viscosity or the Newtonian limit in which the viscosity approaches a constant at very low shear rates, c˙ the shear rate, T is the temperature of the polymer, A1 , A2 , D1 , D2 , n and t∗ are constants of the Table 1 (Osswald [10]). A1 , A2 , D1 and D2 are data-fitted coefficients, n is the power law index in the high shear rate regime, determined by curve fitting and t∗ is the critical stress level at the transition to shear thinning, determined by curve fitting. – Heat capacity = 1700 J/kg/K – Thermal conductivity = 0.173 W/m/K – Pure solvent melting heat = 105 KJ/kg – Solidus and Liquidus Temperature = 420 K • Air – Incompressible air approach is used. – Viscosity. The viscosity of the air is artificially increased to 0.1 Pa • s instead of its value of 1.8 • 10 −5 Pa • s, the reason is that high ratio of viscosity between phases leads to divergences. This strategy is also used by Tofteberg [8]. • Steel. Although in the experimental part a diamond-like carbon coating is used, the thickness of the coating is insignificant and, hence, not taken into account in the simulation parameters.

UN

424

447 448 449 450 451 452 453 454

D

423

the nanocavity, the pressure profile is considered constant through all the inlet. • Walls: No shear rate wall is used. Reason is that using no-slip boundary condition the simulation leads to unphysical results due to the contact line discontinuity of the Navier-Stokes equation, what is already reported in the literature (Gal [11]) and specifically for flows at the nanometer scale (Eijkel [7]). Apart from the theory, Kuhn [12] publishes AFM sections of different short shots at different mould temperatures of moulded v-grooves and it is clearly seen that the contact line went into the grooves, hence, slipping through the wall. View Fig. 6 for a comparison of wall slip models. • Outlet: 1 bar absolute pressure outlet is used. This approach means that the compression of the gas due to the trapping of the air is not taken into account, if the air pressure is calculated using ideal gas law we obtain P = P0 /(1 − x/L), where x is the filled depth of the cavity and L the total depth of the cavity. This equation, once plotted is shown in Fig. 7. It is observed that our approach has an error minor than 10% if the cavity is filled less than its 10%. When more than this percentage is filled, the simulation will overestimate the filling of the cavity, because the calculation would not take into account the opposition done by the air. Future work will be the implementation of a compressible air material and the consideration of a real wall instead of a pressure outlet. From an industrial point of view, since air trap

457

477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515

2.1.4. Boundary conditions

516 517

• Inlet: A transient pressure inlet is used via a Profile function. The profile is extracted from the macro simulation in the node where the nanocavity is. Due to the small size of

518 519 520

455 456

476

521

Table 2 Factorial design.

522 523

458

Factor

Low value

High value

524

459

Tmelt Tmould Filling time Cavity length (L)

250 ◦ C 70 ◦ C 1s 1000 nm

310 ◦ C 100 ◦ C 3s 2000 nm

525

460 461 462

526

Fig. 8. Mould after DLC coating: (a) picture, (b) image of a transverse section using FIB.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

527 528

ARTICLE IN PRESS J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

5

2.1.5. Initial conditions The lower region is patched with a VOF = 1 of polymer, upper region is patched to be air at the same temperature of the walls. Reason for setting the pressure inlet away of the cavity beginning is to take into account the pressure loss of the polymer when getting into the cavity (Tofteberg [8]). Initially, zero velocity at all the model is considered.

529 530 531 532 533 534 535

595 596 597 598 599 600 601

536

602

537

603

539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554

Fig. 9. Designed FIB marks and their position for mould (not to scale). The diameter of the piece is 50 mm and the dimensions of patterns G and P are: A = 4lm, B = 20lm and A = 2lm and B = 10lm, respectively.

D

555 556

563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583

EC

562

RR

561

CO

560

is an effecting factor, injection mould vents should be considered in the design of the mould. Another way to vent a mould is the use of a mould vacuum. This is a very expensive option; however, sometimes it is the only option. This is achieved by sealing all parts of the mould cavity by the use of an O-ring. Once the mould is closed, all of the air is drawn out of the cavity before or at the same time the injection phase of the injection moulding cycle is started. With no air present in the cavity, a perfect condition has been created for trouble-free filling of the cavity (Toth Mold/Die Inc.). • Symmetry: Symmetry is used in order to take into account that there is more fluid at the left and the right of the computationally domain of where we put the fluid (see Fig. 4). The symmetry boundary condition states zero normal velocity at the symmetry line and zero normal gradients of all variables at the symmetry line.

604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629

2.2. Experimental setup

630 631

A mould was specifically designed in order to explore different parameters on the plastic injection process, and facilitate the ulterior characterization of both, the mould and the injected plastic parts, by AFM and SEM. In the mould, specific micro/nano scale patterns have been fabricated with the goal of comparing the replication of such marks in the plastic parts with simulations described in the previous section. Next, we present a detailed description of the analysis performed on the mould and the plastic injection pieces.

UN

559

TE

557 558

2.1.6. Design of experiments A four factors and two levels factorial experiment is carried out. The temperature of the melt, the temperature of the mould and the filling time are process factors, whereas cavity length is a size factor. The levels are shown in Table 2. It is important to state that the filling time determines the needed injection pressure. If the injection time to fill the cavity is small, it must have high injection speed, therefore, must have very high pressure. When the injection time is very long, the melt temperature decreases, viscosity increases, thus it increases the difficulty to fill and requires high injection pressure. There is a lowest point in the middle of the curve and this point is the best filling time which is corresponding to the lowest injection pressure (Fallis [13]). For the used filling times of 1 s and 3 s we see that 1 s needs more injection pressure than 3 s, so the 3 s value is already at the left of the minimum filling time. Concerning to the cavity length, a low level of 1000 nm is used because even with this length, the injections with the low level of mould temperatures and the high level of filling time lead to a zero replication. For the simulations, a maximum injection pressure of 134 MPa is set, which is the maximum injection pressure of the Babyplast 6/10p with a 14 mm screw that IQS University has.

PR OO F

538

632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649

584

650

585

651

586

652

587

653

588

654

589

655

590

656

591

657

592

658

593

659

594

Fig. 10. SEM images of the FIB patterns on mould, as described in Fig. 9.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

660

ARTICLE IN PRESS 6

J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

Diamond Like Carbon [DLC] layer at 170 ◦ C using a CemeCon CC800/9 magnetron sputtering Physical Vapour Deposition (PVD) unit from Cr and graphite targets. The overall coating thickness is 4.75 lm consisting of a gradient adhesion multilayer (Cr + CrN + CrCN + CrC) plus a final amorphous carbon (a-C) layer of 1.8 lm, which exhibits a hydrogen content of approx. 18%. The morphology and roughness of the surfaces were determined with SEM and AFM. The resulting Sq value (≈ 8 nm) is larger as compared to the uncoated surfaces (≈3.5 nm). In addition, the microhardness was determined (3000 HV), using a Fischerscope H100 dynamic microprobe apparatus with a conventional Vickers indenter (load of 10 mN). View Fig. 8 for the mould coating images.

661 662 663 664 665 666 667 668 669 670 671 672

675 676 677 678 679 680 681 682





Fig. 11. Image of the simulation with Tmelt = 250 C, Tmould = 70 C, filling time = 2 s and cavity length is 2000 nm. The H measure is taken at the first time step where all the fluid inside the cavity is already below its glass transition temperature. Notice that the fluid progression is smaller in the vicinity of the walls than in the center of the feature.

683 684

688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703

TE

687

706 707 708

CO

704 705

2.2.2. DLC coating of mould Moulds are usually coated in order to increase their durability and achieve longer lifetimes while reducing maintenance and operational costs. The polished mould was coated at Flubetech with a

711

Table 3 Series of simulations.

712 713 714 715 716 717 718 719 720 721 722 723 724 725 726

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

UN

709 710

2.3. Plastic injection using mould

EC

686

2.2.1. Design and fabrication of mould The mould was designed and fabricated in such a way that: (i) the mould can be easily disassembled, (ii) the surfaces in contact with the injected plastic are flat and can be measured with a commercial AFM instrument (in our case a Bruker ICON) and (iii) plastic can be injected with radial flow. Mould was machined using coated carbide tools from modified AISI 420 steel named Mirrax 40 (by Uddeholm). This steel is a pre-hardened (40 HRC) re-melted stainless tool steel (composition: C 0.21%, Si 0.9%, Mn 0.45%, Cr 13.5%, Mo 0.2%, Ni 0.6%, V 0.25%). The mould was then hardened (1020◦ C, 30 minutes, gas quenched) and tempered (450 ◦ C, 2 h) to a hardness value of 52 HRC. Once the mould was fabricated, the critical surfaces were polished to optical quality finish. Upon polishing, the roughness decreased from approximately Sq = 17 nm to Sq = 3.5 nm, as determined by AFM. This process was carried out using vibrational hand tools following several steps: (i) grinding stone 320–600 (ii) paper 500 (iii) acryl cloth, diamond fluid 15 lm (iv) felt cloth, diamond fluid 3 lm (v) cotton cloth, diamond fluid 1 lm.

RR

685

2.2.3. FIB marks on mould Once the mould was characterized after the DLC coating, different shapes were designed for patterning the surface of the mould by FIB milling. The idea behind this is to extract information about how different variables like roughness, orientation of the patterns, size, shape, etc. can affect and determine the pattern-transfer to the final piece in the injection process. The designed patterns (crosses, rectangles and triangles) are shown in the Fig. 9. The chosen selection was then placed at different distances from the center of the mould. In this way, apart from the effect of the size and shape, we can extract information about how does the fluid flow-velocity affect the pattern transfer, an issue of great relevance in the injection process. A depth of 1 lm was initially selected for all the marks, which was achieved with a 60.000 ls/cm2 dose determined from previous calibration tests on analogous DLC coatings on test samples of the same stainless steel. SEM images of the patterns on the mould at different locations, as indicated in Fig. 9, are shown in Fig. 10. The depth of these marks was 875 ± 25 nm as determined by AFM.

D

674

PR OO F

673

Microinjection moulding was performed using transparent polycarbonate (PC SABIC® LEXAN Resin 133R). Injection was performed in a Babyplast 610P plastic injection machine at a mould temperature of 90 ◦ C and injection pressure of 50 bar with the channel at 280 ◦ C. The pressure was hold for 3 s in a first step and another 4 s for a second pressure of 70 bar.

727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767

3. Results and discussion

768 769

This section is divided in two subsections: one for the simulation results (Section 3.1) and another one for the experimental results (Section 3.2).

770 771 772 773

3.1. Simulation results

774 775

Tmelt /◦ C

Tmould /◦ C

t/s

L/nm

H/nm

250 310 250 310 250 310 250 310 250 310 250 310 250 310 250 310

70 70 100 100 70 70 100 100 70 70 100 100 70 70 100 100

1 1 1 1 3 3 3 3 1 1 1 1 3 3 3 3

1000 1000 1000 1000 1000 1000 1000 1000 2000 2000 2000 2000 2000 2000 2000 2000

5.0 5.0 55.0 55.0 0.0 0.0 55.0 55.0 170.0 420.0 412.3 1500.0 131.5 308.3 350.7 762.5

The replicated height (H) for each simulation is measured as the height reached by the polymer at the center of the cavity when all the fluid inside the cavity is below its glass transition temperature, as it is shown in Fig. 11. It must be pointed out that some authors (Rytka [14] or Mannella [15]) consider that the polymer stops its movement not when its temperature is below the glass transition temperature but when is below its no flow temperature (NFT). The NFT does not have a standard method for measuring it and is considered a not well defined value. A common rule of thumb is to consider NFT = Tg + 30 ◦ C. If we used this approach, all the heights would be slightly smaller. The results for all the simulations are summarized in the Table 3. The analysis of this 24 factorial experiment leads to the Pareto diagram of Fig. 12. The effects are sorted from most significant to least significant and the length of each bar is proportional to the standardized effect, which equals the magnitude of the t-statistic that would be used to test the statistical significance of that effect. A

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792

Q7

ARTICLE IN PRESS J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

7 859

794

860

795

861

796

862

797

863

798

864

799

865

800

866

801

867

802

868

803

869

804

870

805

871

PR OO F

793

806 807 808 809 810 811

818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858

875 876 877

Hattori [21] shows an interesting experimental approach consisting in using a thin-film heater used for heating the mould surface in the nanostructured region, this, in concordance with our simulation results, leads to an improvement of the cavities replication. Rytka [14] concludes that variothermal injection moulding can replicate with high accuracy high aspect ratio microstructures when combined with a polymer of low melt viscosity. Related to the filling time, it is seen that the increase of the filling time leads to a fewer replication of the mark, if we look at the pressure profile of the cavities we obtain the plot of the Fig. 14. We see clearly that with larger filling times, the pressure that reaches the cavity during the time before freezing (which is for all the polymers fewer than the 0.692 s) decreases. The positive effect of the pressure in the replication has already been addressed in the literature by Chen [17] and Pranov [22]. From an industrial point of view it is concluded therefore, that for improving the replication of cavities larger injection pressures are needed, which are obtained only with injection machines with large hydraulic pressures available. Related to the effect of the holding pressure, from a fluid dynamics point of view, the holding pressure has no influence in the cavities replication. Reason is that the holding pressure is applied after the filling stage, and by then, the polymer is already a solid inside the cavities due to the large surface-to-volume ratio. However, divergences are found in the literature when reading experimental papers. Sha [19] shows experimentally that the holding pressure could slightly help to improve the replication, however, for PP and ABS a decrease of the filling length of microcavities was observed. Tofteberg [23] experimentally shows that the holding pressure is an effecting factor, like Sha [19] work, Tofteberg experiments show that

D

817

Fig. 13. Principal effects diagram. It is seen that both the temperatures and the cavity length have a positive effect, in contrast to the filling time negative effect. The cavity length is the factor with a major impact in the cavities replication.

873 874

TE

816

EC

815

RR

814

vertical line is drawn at the location of the 0.05 critical value for Student’s t. Any bars that extend to the right of that line indicate effects that are statistically significant at the 5% significance level. The principal effects diagram is shown in Fig. 13. The effects’ results are summarized in Table 4 The factor with the biggest effect is the geometry of the mark. It is broadly shown in the literature that cavities with small aspect ratio are more difficult to fill, reason is the surface-to-volume ratio leads to an increase of the heat transfer and, hence, to the faster solidification of the polymer (Piotter [16], Chen [17]). We see that both the temperature of the mould and the temperature of the melt are factors that help similarly the replication of the marks. This is a logic result since the marks are copied until the polymer reaches the glass transition temperature, hence, the bigger the mould and melt temperatures, the bigger the time the polymer has to go into the mark before reaching its glass transition temperature. Besides that, the bigger the temperature, the less viscous the polymer is and the easier it goes into the marks. This results are in accordance with experimental works carried out by Kari Mönkkönen [18], Tofteberg and Andreassen [8], Chen [17] and Sha [19]. From an industrial point of view, an increase of these temperatures lead to a more expensive process due to (i) more time is needed for cooling the injected part to the ejection temperature and (ii) the larger energy consumption needed for warming both the mould and the polymer. Concerning to the range where the mould and wall temperatures should be modified, Stormonth [20] comments that the changes in the mould and wall temperature has a bigger effect when set around the optimal temperatures suggested when using conventional injection moulding.

CO

813

UN

812

872

Fig. 12. Pareto diagram. It is observed that the more representative factors are the cavity length and the mould temperature.

878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911

Table 4 Effects table for the replicated height. Effect

Approximated

Average A:Tmelt B:Tmould C:Filling time D:Cavity length AB AC AD BC BD CD

267.8 240.8 275.7 −119.9 478.2 134.1 −93.6 240.8 −79.9 223.2 −117.4

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

912 913 914 915 916 917 918 919 920 921 922 923 924

Q8

ARTICLE IN PRESS 8

J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

about 0.5 s, what prevents further filling of the cavity once in the packing phase. An important success of our model is that we don’t tune the heat transfer coefficient like the simulation approaches of Tofteberg [8], Gang [25] or Rytka [24]. Instead, we let the CFD solver calculate the fluid-side heat transfer based on the local flow-field conditions (e.g., turbulence level, temperature, and velocity profiles), using the following equation in a laminar flow:

925 926 927 928 929 930 931 932 933



934

q = kf

935

∂T ∂n

(3) wall

PR OO F

where n is the local coordinate normal to the wall and kf the thermal conductivity of the fluid.

938 939

Fig. 14. Filling time vs. pressure, data from the macrosimulation. It is observed that, even though the 3s filling time leads to a bigger pressure at advanced times, the 1s filling time leads to a bigger pressure in the beginning of the injection, when the polymer is still fluid. This is the reason for the negative effect of the filling time factor.

944 945

948 949 950 951 952 953 954 955 956 957 958 959 960 961

sometimes the holding pressure has a positive effect and in other experiments has a negative effect. Chen [17] experimentally studied the effect of the holding pressure and shows a positive effect of the holding pressure in the depth of the microchannel and a negative effect of the holding pressure in the width of the microchannel, though of less importance than other factors like the mould temperature, the melt temperature and the filling time. Rytka ([14]) carried out experiments using compression moulding and found no influence on the replicated structure height using an additional compression stroke, reason is, according to their publication, a rapid formation of a frozen polymer layer due to fast cooling upon contact with the mould, this frozen layer prevents further filling of the micro-cavities in the mould during the compression phase. Rytka exposes [24] that the packing phase is less relevant for mould temperatures below the no-flow temperature, because the polymer in the cavity has cooled down below the no flow temperature after

TE

947

962

RR

963 964 965 966

970 971 972 973 974 975 976 977 978 979

UN

969

CO

967 968

The patterns on the injected pieces have been characterized by AFM. The results are shown in Figs. 15 and 16 for positions G2 and P2 (center of the mould), respectively. The marks placed close to the central hole (G3 and P3) disappear during the removal of the perpendicular stick. The marks placed far from the injection point (G1 and P3) are not much replicated, the reason is that these cavities reach smaller pressures and, as proved in the simulations, the pressure is a factor with a positive effect. The most striking observation is the few replication of the cavities, the injected patterns exhibit heights below 100 nm, far from the 900 nm depth from the (negative) mould part. It is observed that the polymer reaches a larger depth when the marks are bigger (G2 compared to P2). This is a result in good correlation with the simulation results, where it was stated that the cavity length is a factor with a positive effect. It is observed in the AFM measurements, that the shape and the depth reached by the polymer in the cross shaped cavities is the same for both directions: parallel and perpendicular to the polymer flow. This experimental observation reinforces the two stages approach used in the simulation part.

EC

946

3.2. Experimental results

D

940

943

993 994 995 996 997 998

1000 1001 1002

937

942

992

999



936

941

991

1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045

980

1046

981

1047

982

1048

983

1049

984

1050

985

1051

986

1052

987

1053

988

1054

989

1055

990

Fig. 15. AFM images of the transferred patterns placed at the center of the mould (G2) and corresponding cross sections.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

1056

ARTICLE IN PRESS J. Pina-Estany et al.

/ International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

1123

1058

1124

1059

1125

1060

1126

1061

1127

1062

1128

1063

1129

1064

1130

1065

1131

1066

1132

1067

1133

1068

1134

1069

1135

RO OF

1057

1070 1071 1072 1073 1074 1075 1076 1077 1078 1079

DP

1080 1081

Fig. 16. AFM images of the transferred patterns placed at the center of the mould (P2) and corresponding cross sections.

1082 1083 1084

1086

TE

1085

References

4. Conclusions

1087

1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110

EC

1090

CO RR

1089

It is concluded that the submodeling approach is a valid approach for simulating how nanocavities are filled in the injection moulding manufacturing process. Using the design of experiments of the Table 3 the main effects are quantified as shown in the Pareto diagram of the Fig. 12. It is concluded that the cavity length, the temperature of the mould and the temperature of the polymer are positive effects with 5, 3 and 2.6 standardized effects values. On the contrary, the filling time is a factor with a negative standardized effect of −1.3. Concerning to the experimental part, it is concluded that the FIB technique is able to manufacture cavities at the nanoscale whose measuring using an AFM gives light in understanding how the polymer flows into the marks. Replication of the marks far from the injection point is poor due to the smaller pressure reached in this area, which is a result in good correlation with the simulation part. Cavity length is experimentally observed to be a factor with a positive effect in the replication. The fact that the shape of the polymer in the cross shaped cavities is the same for both directions (perpendicular and parallel to the flow) reinforces the two stages simulation approach. Future work in the simulation will be the expansion of the algorithm to the three dimensions, this will provide us more know-how related to how the geometry effects the replication of the nanocavities.

UN

1088

1111 1112 1113 1114

Acknowledgments

1115

Q9

1116 1117 1118

Q10

9

1119 1120 1121 1122

This work has been performed within the aim4np project (Automated in-line Metrology for Nanoscale Production), which is supported by the EC through a grant (contract Nr. 309558) within the Seventh Framework Programme NMPCall 2012.1.4-3 on Nanoscale mechanical metrology for industrial processes and products. ICN2 acknowledges support of the Spanish MINECO through the Severo Ochoa Centers of Excellence Program under Grant SEV-2013-0295.

[1] H. Oh, J. Park, Y. Song, J. Youn, Micro-injection moulding of lab-on-a-chip ( LOC ), 19. 2011. [2] Y.E. Yoo, T.H. Kim, D.S. Choi, S.M. Hyun, H.J. Lee, K.H. Lee, S.K. Kim, B.H. Kim, Y.H. Seo, H.G. Lee, J.S. Lee, Injection molding of a nanostructured plate and measurement of its surface properties, Curr. Appl. Phys. 9 (Suppl. 2) (2009) e12–e18. [3] S. Kim, U.T. Jung, S.K. Kim, J.H. Lee, H.S. Choi, C.S. Kim, M.Y. Jeong, Nanostructured multifunctional surface with antireflective and antimicrobial characteristics, ACS Appl. Mater. Interfaces 7 (1) (2015) 326–331. [4] S. Baruah, S.K. Pal, J. Dutta, Nanostructured Zinc Oxide for Water Treatment, Nanosci. Technol.-Asia 2 (2) (2012) 90–102. [5] A.B. Christiansen, J.S. Clausen, N.A. Mortensen, A. Kristensen, Injection moulding antireflective nanostructures, Microelectron. Eng. 121 (2014) 47–50. [6] A.A. Fagade, D.O. Kazmer, Early Cost Estimation for Injection Molded Parts, J. Injection Molding Technol. 3 (4) (2000) 97–106. [7] J.C.T. Eijkel, A. van den Berg, Nanofluidics: What is it and what can we expect from it? Microfluid. Nanofluid. 1 (3) (2005) 249–267. [8] T. Tofteberg, E. Andreassen, Injection moulding of microfeatured parts, Proceedings of the Polymer Processing Society 24th Annual Meeting, PPS-24. 2008, [9] L. Yu, L.J. Lee, K.W. Koelling, Flow and heat transfer simulation of injection molding with microstructures, Polym. Eng. Sci. 44 (10) (2004) 1866–1876. [10] T. Osswald, N. Rudolph, Polymer Rheology, Hanser. 2014. [11] C.G. Gal, M. Grasselli, A. Miranville, Cahn-Hilliard-Navier-Stokes systems with moving contact lines, Calc. Var. 55 (3) (2016) 1–47. [12] S. Kuhn, A. Burr, M. Kübler, M. Deckert, C. Bleesen, Study on the replication quality of micro-structures in the injection molding process with dynamical tool tempering systems, Microsyst. Technol. 16 (10) (2010) 1787–1801. [13] A.G. Fallis, Optimizing the Filling Time and Gate of the Injection Mold on Plastic Air Intake Manifold of Engines, J. Chem. Inf. Model. 53 (9) (2013) 1689–1699. [14] C. Rytka, P.M. Kristiansen, A. Neyer, Iso- and variothermal injection compression moulding of polymer micro- and nanostructures for optical and medical applications, 2015. [15] G.A. Mannella, V. La Carrubba, V. Brucato, W. Zoetelief, G. Haagh, No-flow temperature in injection molding simulation, J. Appl. Polym. Sci. 119 (6) (2011) 3382–3392. [16] V. Piotter, K. Mueller, K. Plewa, R. Ruprecht, J. Hausselt, Performance and simulation of thermoplastic micro injection molding, Microsyst. Technol. 8 (6) (2002) 387–390. [17] C.S. Chen, S.C. Chen, W.H. Liao, R.D. Chien, S.H. Lin, Micro injection molding of a micro-fluidic platform, Int. Commun. Heat Mass Transf. 37 (9) (2010) 1290–1294. [18] K. Monkkonen, T.T. Pakkanen, J. Hietala, E.J. Pakkonen, P. Pakkonen, T. Jskelinen, T. Kaikuranta, Replication of sub-micron features using amorphous thermoplastics, Polym. Eng. Sci. 42 (7) (2002) 1600–1608.

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003

1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164Q11 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188

Q12

ARTICLE IN PRESS 10 1189 1190 1191 1192 1193 1194 1195 1196 1197

J. Pina-Estany et al. / International Communications in Heat and Mass Transfer xxx (2016) xxx–xxx

[19] B. Sha, S. Dimov, C. Griffiths, M.S. Packianather, Micro-injection moulding: Factors affecting the achievable aspect ratios, Int. J. Adv. Manuf. Technol. 33 (1–2) (2007) 147–156. [20] J.M. Stormonth-Darling, Fabrication of difficult nanostructures by injection moulding, 2013, 1–157. [21] S. Hattori, K. Nagato, T. Hamaguchi, M. Nakao, Rapid injection molding of high-aspect-ratio nanostructures, Microelectron. Eng. 87 (5–8) (2010) 1546–1549. [22] H. Pranov, H.K. Rasmussen, N.B. Larsen, N. Gadegaard, On the injection molding of nanostructured polymer surfaces, Polym. Eng. Sci. 46 (2) (2006) 160–171.

[23] T.R. Tofteberg, H. Amédro, F. Grytten, E. Andreassen, Effects of Injection Molding Holding Pressure on the Replication of Surface Microfeatures, Int. Polym. Process. 25 (3) (jul 2010) 236–241. [24] C. Rytka, J. Lungershausen, P.M. Kristiansen, A. Neyer, 3D filling simulation of micro- and nanostructures in comparison to iso- and variothermal injection moulding trials, J. Micromech. Microeng. 26 (6) (2016) 65018. [25] W.W. Kim, M.G. Gang, B.K. Min, W.B. Kim, Experimental and numerical investigations of cavity filling process in injection moulding for microcantilever structures, Int. J. Adv. Manuf. Technol. (2014) 293–304.

1255 1256 1257 1258 1259 1260 1261 1262 1263 1264

1199

1265

1200

1266

1201

1267

PR OO F

1198

1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214

D

1216

TE

1217 1218 1219 1220 1221

EC

1222 1223 1224 1225 1226

RR

1227 1228 1229 1230

1235 1236 1237 1238 1239 1240 1241 1242 1243

UN

1234

CO

1231

1233

1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280

1215

1232

1268

1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309

1244

1310

1245

1311

1246

1312

1247

1313

1248

1314

1249

1315

1250

1316

1251

1317

1252

1318

1253

1319

1254

1320

Please cite this article as: J. Pina-Estany et al., A statistical analysis of nanocavities replication applied to injection moulding, International Communications in Heat and Mass Transfer (2016), http://dx.doi.org/10.1016/j.icheatmasstransfer.2016.11.003