+ All Categories
Home > Documents > Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome...

Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome...

Date post: 15-Aug-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
8
Applied Surface Science 366 (2016) 389–396 Contents lists available at ScienceDirect Applied Surface Science jou rn al h om ep age: www.elsevier.com/locate/apsusc Laser-induced forward transfer of high-viscosity silver pastes D. Munoz-Martin a,, C.F. Brasz b , Y. Chen a , M. Morales a , C.B. Arnold b , C. Molpeceres a a Centro Láser, Universidad Politécnica de Madrid, Alan Turing 1, 28038, Madrid, Spain b Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey, 08544, USA a r t i c l e i n f o Article history: Received 17 November 2015 Received in revised form 21 December 2015 Accepted 5 January 2016 Available online 7 January 2016 Keywords: Laser-induced forward transfer Laser direct-write Printing Metallization a b s t r a c t In this work, a study of the morphology of individual dots of silver paste deposited by laser-induced forward transfer (LIFT) is performed using a ns-pulsed laser at 532 nm. The LIFT process is characterized by scanning confocal microscopy on the deposited material and in-situ time-resolved imaging during the transfer in order to illuminate the flow dynamics in relation to the pulse energy and paste thickness. The influence of process parameters on the structure of transferred dots is explained both phenomenologically and analytically. Depending on the experimental conditions, different transfer regimes were observed. These regimes have similarities to those reported for LIFT of Newtonian fluids and nanopastes, but the multiphase and non-Newtonian rheology and thicker films used lead to noticeable differences, such as the formation of a continuous and stable pillar connecting donor and acceptor substrates when the paste film is thick enough and the energy is optimum. This process regime allows transfer of dots with high aspect ratios, which is desirable for the printing of contacts on solar cells. © 2016 Elsevier B.V. All rights reserved. 1. Introduction Laser-induced forward transfer (LIFT) is a direct-write laser technique, capable of transferring a variety of materials (especially metallic solid materials or material dissolved in an assisting matrix) in different sizes onto a number of substrates [1]. LIFT uses laser pulses to push small volumes of material from a donor film onto an acceptor substrate. The film coats a transparent substrate, and the laser beam is focused at the donor substrate/film interface. Throughout the duration of the pulse, the laser energy is deposited at the laser spot into the interface, vaporizing a small amount of the material, and pushing and accelerating the non-vaporized part of the donor film towards the acceptor substrate [1]. Attractive features of the LIFT process such as easy setup, high flexibility in choice of printed contents and cost-effectiveness have even made an industrial implementation feasible [2]. It is thus possible to use LIFT for printing metallic contacts [3,4] and patterning solder paste [5] for microelectronics, or microwave interconnects in coplanar waveguides [6]. In those cases, inks or pastes with silver nanoparticles are used as donor materials. LIFT can also be used for the metallization of the front side of solar cells [7]. In this case, one of the aims of solar cell researchers and manufacturers is to find technologies leading to an increase in the Corresponding author. Tel.: +34 913365540. E-mail address: [email protected] (D. Munoz-Martin). efficiencies of solar cells while keeping costs low. Specifically, pro- cedures capable of making better contacts by improving the aspect ratio and decreasing contact losses are sought [8]. Several different techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and robust method to deposit metal contacts in solar cell industrial production. LIFT performed with ns pulses is a well-known tech- nology to generate structured metallization onto substrates, but it has not yet been used to create the front contact fingers for a pho- tovoltaic device in a single step, although a two-step approach has been developed recently [10]. Previous work [7] demonstrated the feasibility of printing long, high-aspect-ratio lines of a commercial micron-sized silver paste onto c-Si cells, although the parametric window to achieve good results was narrow. In order to control this printing process, it is desirable to understand the effects of the LIFT process parameters (laser pulse energy, silver paste donor thickness, and gap distance between donor and acceptor substrates) on the morphology of the transferred paste. The effect of these experimental parameters on the mechanisms for material transfer using LIFT have been widely discussed in the case of solids [11], Newtonian fluids [12–14], and nanopastes [15–17]. However, the particular rheology, viscosity (>200 Pa s), and particle size of the silver paste (1–10 m) used strongly affect the transfer process and the shape and quality of the printed voxel, distinguishing its mechanisms from those observed previously in conductive inks or nanopastes. The present work employs single Gaussian laser pulses with spot size on the order http://dx.doi.org/10.1016/j.apsusc.2016.01.029 0169-4332/© 2016 Elsevier B.V. All rights reserved.
Transcript
Page 1: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

L

Da

b

a

ARR2AA

KLLPM

1

tmipatTatofca

aipccm

h0

Applied Surface Science 366 (2016) 389–396

Contents lists available at ScienceDirect

Applied Surface Science

jou rn al h om ep age: www.elsev ier .com/ locate /apsusc

aser-induced forward transfer of high-viscosity silver pastes

. Munoz-Martina,∗, C.F. Braszb, Y. Chena, M. Moralesa, C.B. Arnoldb, C. Molpeceresa

Centro Láser, Universidad Politécnica de Madrid, Alan Turing 1, 28038, Madrid, SpainDepartment of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey, 08544, USA

r t i c l e i n f o

rticle history:eceived 17 November 2015eceived in revised form1 December 2015ccepted 5 January 2016vailable online 7 January 2016

eywords:

a b s t r a c t

In this work, a study of the morphology of individual dots of silver paste deposited by laser-inducedforward transfer (LIFT) is performed using a ns-pulsed laser at 532 nm. The LIFT process is characterizedby scanning confocal microscopy on the deposited material and in-situ time-resolved imaging during thetransfer in order to illuminate the flow dynamics in relation to the pulse energy and paste thickness. Theinfluence of process parameters on the structure of transferred dots is explained both phenomenologicallyand analytically.

Depending on the experimental conditions, different transfer regimes were observed. These regimes

aser-induced forward transferaser direct-writerintingetallization

have similarities to those reported for LIFT of Newtonian fluids and nanopastes, but the multiphase andnon-Newtonian rheology and thicker films used lead to noticeable differences, such as the formationof a continuous and stable pillar connecting donor and acceptor substrates when the paste film is thickenough and the energy is optimum. This process regime allows transfer of dots with high aspect ratios,which is desirable for the printing of contacts on solar cells.

. Introduction

Laser-induced forward transfer (LIFT) is a direct-write laserechnique, capable of transferring a variety of materials (especially

etallic solid materials or material dissolved in an assisting matrix)n different sizes onto a number of substrates [1]. LIFT uses laserulses to push small volumes of material from a donor film onton acceptor substrate. The film coats a transparent substrate, andhe laser beam is focused at the donor substrate/film interface.hroughout the duration of the pulse, the laser energy is depositedt the laser spot into the interface, vaporizing a small amount ofhe material, and pushing and accelerating the non-vaporized partf the donor film towards the acceptor substrate [1]. Attractiveeatures of the LIFT process such as easy setup, high flexibility inhoice of printed contents and cost-effectiveness have even maden industrial implementation feasible [2].

It is thus possible to use LIFT for printing metallic contacts [3,4]nd patterning solder paste [5] for microelectronics, or microwaventerconnects in coplanar waveguides [6]. In those cases, inks orastes with silver nanoparticles are used as donor materials. LIFT

an also be used for the metallization of the front side of solarells [7]. In this case, one of the aims of solar cell researchers andanufacturers is to find technologies leading to an increase in the

∗ Corresponding author. Tel.: +34 913365540.E-mail address: [email protected] (D. Munoz-Martin).

ttp://dx.doi.org/10.1016/j.apsusc.2016.01.029169-4332/© 2016 Elsevier B.V. All rights reserved.

© 2016 Elsevier B.V. All rights reserved.

efficiencies of solar cells while keeping costs low. Specifically, pro-cedures capable of making better contacts by improving the aspectratio and decreasing contact losses are sought [8]. Several differenttechniques [8,9] have been developed to overcome the limitationsof conventional screen-printing, which is the well-established androbust method to deposit metal contacts in solar cell industrialproduction. LIFT performed with ns pulses is a well-known tech-nology to generate structured metallization onto substrates, but ithas not yet been used to create the front contact fingers for a pho-tovoltaic device in a single step, although a two-step approach hasbeen developed recently [10].

Previous work [7] demonstrated the feasibility of printing long,high-aspect-ratio lines of a commercial micron-sized silver pasteonto c-Si cells, although the parametric window to achieve goodresults was narrow. In order to control this printing process, it isdesirable to understand the effects of the LIFT process parameters(laser pulse energy, silver paste donor thickness, and gap distancebetween donor and acceptor substrates) on the morphology of thetransferred paste. The effect of these experimental parameters onthe mechanisms for material transfer using LIFT have been widelydiscussed in the case of solids [11], Newtonian fluids [12–14], andnanopastes [15–17]. However, the particular rheology, viscosity(>200 Pa s), and particle size of the silver paste (∼1–10 m) used

strongly affect the transfer process and the shape and quality of theprinted voxel, distinguishing its mechanisms from those observedpreviously in conductive inks or nanopastes. The present workemploys single Gaussian laser pulses with spot size on the order
Page 2: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

390 D. Munoz-Martin et al. / Applied Surfa

Table 1Dupont Solamet PV17F properties [19].

Viscosity (Brookfield HBT, 10 rpmSC4-14/6R utility cap and spindle,25 C)

280–400 Pa s

Solid Content at 750 C 89.5–91.0%Resistivity <5 m/sq/mSilver grain 1–4 mOrganic carrier N,N′-Ethane-1,2-diylbis(decanamide)

12-Hydroxy-N-[2-[1-oxydecyl)amino]ethyl]octadecanamideN,N′-Ethane-1,2-

oTt[d

2

2

ttlsetm

2tvaHit5pd1nmusmuldfshwdm

elAsitt

diylbis(12-hydroxyoctadecanamide)Thinner 9450

f the donor film thickness to print voxels with high aspect ratios.he morphology of the printed voxel is discussed in terms of theransfer mechanisms. To this end, a time-resolved imaging study14,18] is presented to enable a detailed explanation of the transferynamics for this important class of material.

. Experimental

.1. Voxel printing using LIFT

A commercial silver paste (DuPont Solamet PV17F) was used ashe donor material for LIFT. It is a highly conductive silver pastehat provides excellent efficiency, reliable soldered adhesion, loway down, rapid dry, and very fast firing. Although it is designed forcreen-printing, its excellent electrical and curing/sintering prop-rties make it a good candidate for LIFT metallization. Table 1 showshe physical and chemical properties of the paste provided by the

anufacturer [19].According to PV17F paste datasheet, it has a viscosity between

80 and 400 Pa s. Due to the large influence of the viscosity onhe quality of the printing line resolution, the exact viscosityalue of the paste was measured before every experiment with

cone spindle & plate viscometer (Wells/Brookfiel ViscometerBDV-II; Cone spindle CPE-52). The reference value of the viscos-

ty was 250 ± 30 Pa s. As the silver paste exhibits non-Newtonianhixotropic fluid behavior, the silver paste was gently stirred for

minutes to attain the equilibrium viscosity before using it. Theaste was then deposited onto microscope slides, which act asonor substrates, using a commercial coater (Control Coater model01, RK PrintCoat Instruments Ltd). Two different donor thick-esses were used: 30 ± 5 m and 50 ± 10 m. The thickness waseasured in each donor sample before and after LIFT experiments

sing confocal microscopy. c-Si wafers were used as acceptor sub-trates. The donor substrate was placed at a gap distance of 50 m,easured from the surface of the donor film to acceptor substrate,

sing Kapton tape. A diode pumped, ns-pulsed, solid state Nd:YVO4aser (Spectra Physics Explorer), emitting at 532 nm with a pulseuration of 15 ns was used. The beam was focused into the inter-ace between glass and silver paste to initiate transfers. The opticalystem comprises an attenuator and an optical scanner (ScanlaburrySCAN II) with an f-theta lens (250 mm focal length). The beamaist at the focus is 12.5 m. This value has been measured bothirectly using a beam profiler (Ophir Spiricon) and indirectly byeans of the Liu method [20].In order to study the effect of the different experimental param-

ters, single dots of paste or voxels were transferred using a singleaser shot. Laser peak fluences were varied from 0.6 to 14.4 J/cm2.fter laser irradiation, the donor was removed from the acceptor

ubstrate, leaving the transferred voxels. Fig. 1 shows microscopemages of voxels transferred using four different fluences and thewo donor thicknesses. When using the lowest fluence (0.6 J/cm2),here is no transfer for the thicker donor (50 m), while the

ce Science 366 (2016) 389–396

thinner donor (30 m) leads to small clusters of paste on theacceptor substrate. This suggests the existence of an energy trans-fer threshold that depends on the donor thickness, with a largerthreshold for the 50-m-thick film. This is consistent with pre-vious studies of LIFT using Newtonian fluids [13,21,22], and canbe understood by the observation that more material must beaccelerated in the case of a thicker donor, requiring more energy.When the pulse energy increases there is paste transfer using boththicknesses, although in the case of the 30-m-thick donor filmsthe voxel always exhibits splashing, with finer and more numer-ous droplets at higher fluences. The 50-m-thick donor leads toconcrete-dot transfer, with a single paste dot that shows no splash-ing or disaggregation of the voxel. The size of the dot increasesas energy increases. At the largest fluence (5.9 J/cm2), the dot issurrounded by clearer debris that could be due to cured organicresidue.

The morphology of the voxels on the acceptor substrate and theholes left in the film on the donor substrate were characterizedusing confocal microscopy (Leica DCM 3D). Fig. 2 shows the falsecolor height map of both donor (Fig. 2a) and acceptor (Fig. 2c) sub-strates for the two thicknesses. It also shows the correspondingcross-section profiles (Fig. 2b). The dashed horizontal line in theprofiles indicates the interface between the silver paste and theglass slide.

For the donor films, the holes reached the glass slide, with nosilver paste remaining in the bottom of the hole in either case. Thehole left in the thicker donor is smaller in diameter and the insidewall is steeper compared to the thinner-donor case. A crustal annu-lar bulge that was formed surrounding the crater is observed in bothcases. Furthermore, the width and the shape of the holes are con-gruent with those of the transferred voxels. Widths of holes andvoxels are similar and on the order of 100 m, much larger thanthe laser beam spot. However, there is a large difference in theheight of the voxels and the depth of the hole, which indicates thata significant portion of the paste affected by the laser pulse is nottransferred but displaced to the border of the hole. Changes in thedensity of the paste in the surrounding regions due to the laser irra-diation could also contribute to the large differences in the volumesof the hole and the voxel.

Fig. 3 shows the voxel size (height, width and aspect ratio) as afunction of the laser peak fluence. Since the aim of this work is touse LIFT to metalize devices, the aspect ratio, calculated as the ratioof height to width, can be considered a suitable figure-of-merit ofthe process: large heights result in lower electrical resistances oftransferred lines, while lower widths lead to a smaller area occu-pied by the line in the final device. The width of the voxel is definedas the diameter of the compact circular-like region excluding thesurrounding scattered clusters, and the height is defined as thehighest value of the transversal profile. Each data point in Fig. 3 rep-resents the average of 10 measurements, with error bars denotingone standard deviation. The width of the dots is difficult to definein the low-energy regime for the 30-m-thick donor (between 0.6and 2.0 J/cm2), where the dots are made up of large clumps of paste(Fig. 1a). In the case of the 30-m-thick donor, the height of the vox-els remains almost constant as laser fluence increases, while in thecase of the 50-m-thick donor, the height decreases with energy.Using both donor thicknesses, the width of the voxels increases as afunction of laser fluence in a range between 0.6 and 7.4 J/cm2. Thus,the decrease of aspect ratio with fluence is revealed as a commontrend. Using higher laser fluences (beyond 8 J/cm2), which are notshown, the voxel size has a downward trend and finally saturates.This phenomenon could be due to the absorption of the focused

pulse in the glass substrate exceeding the damage threshold ofglass. In any case, this fluence regime is not suitable for transfer-ring voxels with large aspect ratios, so it was not considered anyfurther. The optimal transferred voxel with the largest aspect ratio
Page 3: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

D. Munoz-Martin et al. / Applied Surface Science 366 (2016) 389–396 391

Fig. 1. Microscope images of LIFT dots transferred using single shots of a ns laser with fluence varying from 0.6 to 5.9 J/cm2 with two different donor thicknesses: (a) 30 mand (b) 50 m.

Fig. 2. Confocal microscope images of (I-a, II-a) donor films and (I-c, II-c) transferreddots from 1.3 J/cm2 laser pulses with films of thickness (I) 30 m and (II) 50 m.Ps

(fl

tl

• just above the transfer threshold (0.6 J/m2). Thus, this is the

rofiles of both the hole left in the donor and the voxel transferred to the acceptorubstrate are shown in (I-b, II-b).

0.13) is obtained using the 50-m-thick donor and the thresholduence for transfer.

To summarize, the experimental results on voxel transfer showhat four different regimes can be considered depending on theaser peak fluence and the donor thickness:

No transfer. When using laser fluences below a threshold valuethat depends on the donor thickness (0.6 and 3.2 J/cm2 for the 30and 50 m thick films respectively), no transfer is observed.Cluster transfer. In the thin-donor case at low fluences(F ≤ 2 J/m2), the voxel consists of non-continuous clusters ofpaste. The voxel width increases as fluence increases, while theheight remains small.Explosive transfer. At high laser fluences (F ≥ 2 J/m2), the transfer

resembles a splash, with larger and relatively uniform clus-ters of paste in the center of the spot and a large quantityof small droplets of paste further from the center. The aspect

Fig. 3. Height, width and aspect ratio of the transferred dots as a function of laserpeak fluence with respect to two different paste thicknesses (30 m and 50 m).

ratio is very small (small height and large width), smaller than0.05.

• Concrete-dot transfer. In the thick-donor case, a single, well-defined dot of paste is observed for the whole range of fluencesstudied. As the laser fluence increases, the height decreases andthe width increases, leading to large aspect ratios for fluences

mechanism of technological interest, because it allows success-ful transfer of large aspect ratio and uniform lines of paste usingLIFT.

Page 4: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

3 Surfa

2

tLewNddpoWIll

lcm3tlitaate

2

fsflsmfttefdreib

Fdetsi[fm8aiiic

ai

92 D. Munoz-Martin et al. / Applied

.2. Time-resolved imaging experiment

To gain insight into the transfer mechanisms responsible forhe various regimes observed earlier, side-view images of theIFT process were obtained with time-resolved shadowgraphy. Thexperimental setup used is similar to that of the previous section,ith differences highlighted here. Pulses from a frequency-tripledd:YVO4 laser (Coherent AVIA) emitting at 355 nm with a pulseuration of 20 ns are focused by a 10x laser objective (NA = 0.25)own to a beam waist of about 15 m to initiate transfer of silveraste. Time-resolved images of the transfer from a side view arebtained with a microscope (InfiniTube with a Mitutoyo 50x Long-D Objective) attached to a CCD camera (SPOT Insight IN1820).

llumination is provided by strobing a 25-ns plasma-discharge flashamp (HSPS Nanolite) after a specified time delay relative to theaser pulse, set by a digital delay generator (SRS DG535) [18].

Although the experimental system used has a different wave-ength and spot size, the objective is to find similar transferonditions to those shown in Section 2.1. Silver paste used as donoraterial is absorbent in the two wavelengths studied (532 and

55 nm). Therefore, the physics underlying the process should behe same in both cases, apart from minor differences in the particu-ar fluence thresholds. In Section 2.2.1, the acceptor substrate is notncluded in order to make illumination easier, following previousime-resolved imaging studies [14,22]. Later, in Section 2.2.2, ancceptor is included to investigate its effect on transfer. The donorssembly is moved with a translation stage after each pulse, and aop-down camera (EO-5012C) is used to ensure pulses are fired farnough away from previous holes in the donor film.

.2.1. Effects of film thickness and pulse energySequences of images at increasing time delays are presented

or various donor thicknesses and laser peak fluences. Fig. 4 showsequences for a silver-paste donor of thickness 120 m at two laseruences. It is immediately clear that the transfer of silver paste isignificantly different from that of Newtonian liquids observed inost time-resolved imaging studies of LIFT [14,22,23]; rather than

orming a smooth jet that breaks up into droplets, the action ofhe laser pulse on the silver paste forms a coherent blister shapehat then fragments into clumps of paste. At the laser peak flu-nce of 8.5 J/cm2 used in Fig. 4a, the protrusion fragments into aew clumps, with a larger primary clump propelled away from theonor. Increasing the peak fluence to 23 J/cm2, the transfer in Fig. 4besembles an explosion, with a more uniform distribution of clumpsxpanding outward over time. In both cases, a crown-like features observed above which the displaced paste remains intact, andelow which fragmentation occurs.

Similar mechanisms of transfer are observed for thinner donors.ig. 5 shows sequences of images for three different energies for aonor with a thickness of approximately 60 m. The lowest flu-nce (2.8 J/cm2) is below the threshold for transfer, resulting inhe formation of a stable blister, shown in Fig. 5a. This behavior isimilar to the blister formation achieved by focusing a laser pulsento a solid polymer film [18,24] and used for blister-actuated LIFT15,25]. Above the transfer threshold, a regime of transferring aew clumps is again observed for 4.2 J/cm2 pulses in Fig. 5b, and

ore explosive transfers are observed at the higher peak fluence of.5 J/cm2 in Fig. 5c. Interestingly, these laser fluences are about halfs large as the corresponding values generating the same behaviorn the thick-donor case (120 m), and the donor thickness (60 m)s also half that of the thick-donor case. The time scales for transfern the thin-donor case are also roughly half that of the thick-donor

ase.

Finally, Fig. 6 shows sequences of images for three energies withn even thinner donor of thickness 12 m. The laser fluences lead-ng to transfer are much lower (0.6 and 2.3 J/cm2) than for thicker

ce Science 366 (2016) 389–396

donors, and the time scale of the whole process is shorter as well.In this thin-donor case, even the lowest fluence imaged, 0.6 J/cm2,leads to explosive transfers. The transition from no transfer to a fewfragmented clumps to a porous explosion appears to occur over asmaller fluence range with thinner donors.

The observation of transfers with a small number of clumpsjust above the threshold fluence and more uniform explosions forhigher fluences correlates well with the transfer regimes of clusterdot transfer and explosion dot transfer observed earlier, respec-tively. To observe the concrete-dot transfer regime, an acceptorsubstrate must be included in the imaging setup.

2.2.2. Influence of an acceptor substrateThe presence of an acceptor substrate can dramatically affect the

transfer, depending on the spacing between the donor and accep-tor. For large enough gaps, the transfers will be similar to thoseimaged in Section 2.2.1, but as the gap gets small enough and of theorder of the donor thickness, confinement can play a large role.

Fig. 7 shows sequences of images taken with an acceptor sub-strate approximately 140 m from the donor film for two donorthicknesses and similar laser fluences. First, in Fig. 7a, transfersfrom a donor of thickness 35 m with a peak fluence of 6.8 J/cm2

are shown to reach the acceptor while exploding, leaving a thinresulting deposit which has splashed to cover a large area. Fig. 7bshows that for the thicker 60-m donor with 8.4 J/cm2 transfers,the confinement of the acceptor prevents the paste from exploding,instead forming a stable, connected bridge akin to liquid bridges.The final shape of the voxel is therefore not obtained until the donorsubstrate is removed from the acceptor.

The donor thickness and peak fluence of Fig. 7b are equivalentto the values used to obtain the exploding transfers in Fig. 5c, so itis clear that the acceptor substrate plays a crucial role in preventingfragmentation. The donor-thickness dependence is consistent withthe trend observed in Section 2.2.1, that pulses of a given energy aremore likely to cause explosive transfers as the donor gets thinner.

3. Discussion

The voxel transfer and imaging results can be understood interms of the effects of the experimental parameters on the inter-nal dynamics of the paste, similar to previous works on Newtonianfluids [13,22,26]. As described above, four transfer regimes havebeen observed, both in the morphology of the transferred voxelsand in the time-resolved images: no transfer, cluster-dot trans-fer, explosive transfer, and concrete-dot transfer. Three importantexperimental parameters considered here are the laser peak flu-ence (F), donor thickness (h), and gap distance (d). Fig. 8 shows aschematic of the LIFT process according to the results shown above.First, the incident laser pulse is focused onto the glass/paste inter-face, rapidly heating and vaporizing the organic component of thesilver paste, forming a high-pressure and high-temperature vaporbubble [22]. For a given donor thickness (h1), smaller than d, it ispossible to define a threshold laser peak fluence (Fth = Fth(h)) fortransfer of material [13]. The energy threshold of the silver paste isattributable to the internal and external force balance. Using pulseenergies below Fth, the bubble grows to only a modest size beforecollapsing, which results in the formation of a blister in the paste(Fig. 8a). Above Fth, the bubble expands to a larger size due to thelarger internal pressure overcoming the ambient air pressure andsurface tension force (Fig. 8b). The deceleration of the bubble leadsto a low-pressure region where the paste protrudes, between the

bubble and the paste-air interface, so the silver paste surroundingthe bubble flows toward this low-pressure region. The flow andstretching of this protrusion of paste eventually leads to fragmen-tation into clusters of paste that are transferred to the acceptor
Page 5: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

D. Munoz-Martin et al. / Applied Surface Science 366 (2016) 389–396 393

Fig. 4. Time-resolved images of LIFT of silver paste from a donor film of thickness 120 m using laser fluences of approximately (a) 8.5 and (b) 23 J/cm2.

Fig. 5. Time-resolved images of LIFT of silver paste from a film of thickness 60 m using laser fluences of approximately (a) 2.8, (b) 4.2, and (c) 8.5 J/cm2.

Page 6: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

394 D. Munoz-Martin et al. / Applied Surface Science 366 (2016) 389–396

Fig. 6. Time-resolved images of LIFT of silver paste from a film of thickness 12 m using laser fluences of approximately (a) 0.6, (b) 2.3, and (b) 8.5 J/cm2.

F donora

set(vfepf

ig. 7. Time-resolved images of LIFT of silver paste with a gap of 140 m betweennd (b) 60 m and 8.4 J/cm2.

ubstrate. For fluences that are close to Fth, most of the kineticnergy in the paste goes into separating these clusters, so thathere are relatively few of them and they travel at low velocitiessee Figs.4a and 5b). For higher fluences, larger and more violentapor bubbles are produced, which results in more kinetic energy

or fragmentation. The radially outward motion induced by bubblexpansion leads to fragmentation into a larger number of smallerieces that move faster, exploding outwards. The subsequent trans-er contains finer droplets and has a larger diameter along with the

and acceptor. The film thicknesses and laser fluences are (a) 35 m and 6.8 J/cm2

hole in the donor substrate. The remaining silver paste recoil formsthe crustal annular bulge of the crater left on the donor film as seenin Fig. 2.

The progression from being below the threshold for trans-fer to cluster-dot transfer and eventually explosive transfer as

laser fluence increases is similar to that of LIFT with Newtonianliquids, where the most confined and reproducible transfers areobtained just above the threshold and splashing is observed forhigher fluences [14]. The distinction between a small number
Page 7: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

D. Munoz-Martin et al. / Applied Surface Science 366 (2016) 389–396 395

Fig. 8. Schematic of transfer mechanisms for varying laser peak fluences and donor thicknesses with a fixed gap distance: (a) blister formation, when the fluence is belowthe transfer threshold; (b) cluster-dot transfer, for fluences slightly above the threshold; (c) explosion-dot transfer, when the fluence is much higher than the threshold; and(d) concrete-dot transfer, when the donor is thick enough relative to the gap that the acceptor prevents fragmentation during transfer. The bottom image in each columns

ocf

ntctmpctTciesba

ptttw

wmdima

hows the final voxel obtained after removing the donor.

f relatively large clusters at lower fluences and explosions of finerlusters at high fluences (Figs.1a, 4 and 5) is a new result that arisesrom the multiphase rheology of the silver paste, however.

Fig. 7b directly demonstrates that a different transfer mecha-ism occurs when the donor thickness h is large enough relative tohe gap d that the protruding paste pillar reaches the acceptor whileonnected to the donor substrate. Once the pillar front impactshe acceptor substrate, confinement prevents the paste from frag-

enting into pieces, stopping the expansion and stretching of theaste being transferred. Eventually, the laser-induced vapor bubbleollapses to an equilibrium shape, leaving a concave pillar takinghe form of a liquid bridge wetting both the donor and acceptor.his pillar is stable due to the high viscosity and non-Newtonianharacter of the paste, inhibiting the spreading and thinning of thempacting pillar and effectively solidifying it. By contrast, in LIFTxperiments with Newtonian liquids, jets that impact an acceptorubstrate are unstable and break up, with some liquid retractingack to the donor film and the rest transferring to a droplet on thecceptor [26].

When the donor substrate is pulled off from the acceptor, theillar breaks, leaving a single concrete-dot of paste on the accep-or substrate. This mechanism explains the lack of splashing in theransfers in Fig. 1b, where the 50-m film is thick enough relativeo the 50-m gap distance so that the pillar impacts the acceptorithout breaking up.

This effect was suggested by the authors [7] when printing linesith heights larger than the gap distance. Previous LIFT experi-ents with nanoparticle inks also revealed spikes in the donor film

ue to remnants of collapsing filaments in the case of nanoparticlenks [15], but this is, to our knowledge, the first imaging experi-

ent in which the formation of stable columns between donor andcceptor substrates is shown.

4. Conclusions

In this work, a study of laser-induced forward transfer (LIFT) ofsingle dots of silver paste was performed using ns-pulsed lasers. Thepaste contains micron-sized particles, and relatively thick donorswere used in order to print voxels with high aspect ratios, leading totransfer mechanisms differing from those previously seen using sil-ver nanopastes or Newtonian fluids. These transfer mechanisms arestudied through the morphology of transferred paste dots obtainedwith confocal microscopy and by time-resolved imaging of thetransfer process. Four transfer regimes are observed and defined:non-dot transfer, when the laser fluence is below the thickness-dependent transfer threshold; cluster-dot transfer, in which thetransferred paste fragments into a small number of large clustersand the dot size increases with laser fluence; concrete-dot trans-fer, where the paste is thick enough and the gap is small enoughto allow the protruding jet to touch the acceptor substrate beforefragmentation; and explosive transfer, where high laser pulse ener-gies cause a bursting transfer of paste that splashes on the acceptor.Concrete-dot transfer is the most useful for printing high-aspect-ratio dots, and the highest aspect ratios were achieved with pulseenergies just above the threshold for transfer.

Acknowledgments

This work was supported by EUROPEAN COMISSION APPOLOFP7-2013-NMP-ICT-FOF.609355 and Spanish MINECO projectsSIMLASPV-MET (ENE2014-58454-R) and HELLO (ENE2013-48629-

C4-3-R). C.F.B. and C.B.A. acknowledge support from the NSF MRSECprogram through the Princeton Center for Complex Materials(grant DMR-0819860) and the Rutgers-Princeton NSF-IGERT onNanotechnology for Clean Energy.
Page 8: Applied Surface Science - Princeton University · techniques [8,9] have been developed to overcome the limitations of conventional screen-printing, which is the well-established and

3 Surfa

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

96 D. Munoz-Martin et al. / Applied

eferences

[1] C.B. Arnold, P. Serra, A. Piqué, Laser direct-write techniques for printing ofcomplex materials, MRS Bull. 32 (2007) 23–32, http://dx.doi.org/10.1557/mrs2007.11.

[2] G. Hennig, Lasersonic® LIFT process for large area digital printing, JLMN 7(2012) 299–305, http://dx.doi.org/10.2961/jlmn.2012.03.0012.

[3] J.M. Fernández-Pradas, P. Serra, Conductive silver ink printing through thelaser-induced forward transfer technique, Appl. Surf. Sci. 255 (2015)5342–5345, http://dx.doi.org/10.1016/j.apsusc.2014.12.100.

[4] A. Piqué, H. Kim, R.C.Y. Auyeung, A.T. Smith, Laser forward transfer offunctional materials for digital fabrication of microelectronics, J. Imaging Sci.Technol. 57 (2013) 1–8, http://dx.doi.org/10.2352/j.imagingsci.technol.2013.57.4.040404.

[5] S.A. Mathews, N.A. Charipar, R.C. Auyeung, H. Kim, A. Piqué, Laser forwardtransfer of solder paste for microelectronics fabrication, in: Proc. SPIE 9351Laser-based Micro- and Nanoprocessing IX, 93510Y, 2015, http://dx.doi.org/10.1117/12.2080410.

[6] E. Breckenfeld, H. Kim, R.C.Y. Auyeung, N. Charipar, P. Serra, Laser-inducedforward transfer of silver nanopaste for microwave interconnects, Appl. Surf.Sci. 331 (2015) 254–261, http://dx.doi.org/10.1016/j.apsusc.2015.01.079.

[7] M. Morales, Y. Chen, D. Munoz-Martin, S. Lauzurica, C. Molpeceres, Highvolume transfer of high viscosity silver pastes using laser direct-writeprocessing for screen printing of c-Si cells, in: Proc. SPIE 9351 Laser-basedMicro- and Nanoprocessing IX, 93510B, 2015, http://dx.doi.org/10.1117/12.2080092.

[8] P.S. Aakella, S. Saravanan, S.S. Joshi, C.S. Solanki, Pre-metallization processesfor c-Si solar cells, Sol. Energy 97 (2013) 388–397, http://dx.doi.org/10.1016/j.solener.2013.08.026.

[9] A. Mette, P.L. Richter, M. Hörteis, S.W. Glunz, Metal aerosol jet printing forsolar cell metallization, Prog. Photovolt. Res. Appl. 15 (2007) 621–627, http://dx.doi.org/10.1002/pip.759.

10] T.C. Roder, E. Hoffmann, J.R. Kohler, J.H. Werner, 30 m wide contacts onsilicon cells by laser transfer, in: 2010 35Th IEEE Photovoltaic SpecialistsConference (PVSC), 2010, pp. 003597–003599, http://dx.doi.org/10.1109/PVSC.2010.5614378.

11] D.P. Banks, C. Grivas, I. Zergioti, R.W. Eason, Ballistic laser-assisted solid

transfer (BLAST) from a thin film precursor, Opt. Express 16 (2008)3249–3257, http://dx.doi.org/10.1364/OE.16.003249.

12] M.S. Brown, C.F. Brasz, Y. Ventikos, C.B. Arnold, Impulsively actuated jets fromthin liquid films for high-resolution printing applications, J. Fluid Mech. 709(2012) 341–370, http://dx.doi.org/10.1017/jfm.2012.337.

[

ce Science 366 (2016) 389–396

13] P. Serra, M. Duocastella, J.M. Fernández-Pradas, J.L. Morenza, Liquidsmicroprinting through laser-induced forward transfer, Appl. Surf. Sci. 255(2009) 5342–5345, http://dx.doi.org/10.1016/j.apsusc.2008.07.200.

14] M. Duocastella, J.M. Fernández-Pradas, J.L. Morenza, P. Serra, Time-resolvedimaging of the laser forward transfer of liquids, J. Appl. Phys. 106 (2009)084907-7, http://dx.doi.org/10.1063/1.3248304.

15] M. Duocastella, H. Kim, P. Serra, A. Piqué, Optimization of laser printing ofnanoparticle suspensions for microelectronic applications, Appl. Phys. A 106(2012) 471–478, http://dx.doi.org/10.1007/s00339-011-6751-z.

16] S.A. Mathews, R.C.Y. Auyeung, H. Kim, N.A. Charipar, A. Piqué, High-speedvideo study of laser-induced forward transfer of silver nano-suspensions, J.Appl. Phys. 114 (2013) 064910-10, http://dx.doi.org/10.1063/1.4817494.

17] C. Boutopoulos, A.P. Alloncle, I. Zergioti, P. Delaporte, A time-resolvedshadowgraphic study of laser transfer of silver nanoparticle ink, Appl. Surf.Sci. 278 (2013) 71–76, http://dx.doi.org/10.1016/j.apsusc.2012.12.002.

18] M.S. Brown, N.T. Kattamis, C.B. Arnold, Time-resolved study of polyimideabsorption layers for blister-actuated laser-induced forward transfer, J. Appl.Phys. 107 (2010) 083103, http://dx.doi.org/10.1063/1.3327432.

19] DuPont, Photovoltaic metallization pastes, http://www.dupont.com/products-and-services/solar-photovoltaic-materials/photovoltaic-metallization-pastes.html.

20] J.M. Liu, Simple technique for measurements of pulsed Gaussian-beam spotsizes, Opt. Lett. OL 7 (1982) 196–198, http://dx.doi.org/10.1364/OL.7.000196.

21] V. Dinca, M. Farsari, D. Kafetzopoulos, A. Popescu, M. Dinescu, C. Fotakis,Patterning parameters for biomolecules microarrays constructed withnanosecond and femtosecond UV lasers, Thin Solid Films 516 (2008)6504–6511, http://dx.doi.org/10.1016/j.tsf.2008.02.043.

22] M.S. Brown, N.T. Kattamis, C.B. Arnold, Time-resolved dynamics oflaser-induced micro-jets from thin liquid films, Microfluid. Nanofluid. 11(2011) 199–207, http://dx.doi.org/10.1007/s10404-011-0787-4.

23] J. Yan, Y. Huang, C. Xu, D.B. Chrisey, Effects of fluid properties and laserfluence on jet formation during laser direct writing of glycerol solution, J.Appl. Phys. 112 (2012) 083105-10, http://dx.doi.org/10.1063/1.4759344.

24] N.T. Kattamis, M.S. Brown, C.B. Arnold, Finite element analysis of blisterformation in laser-induced forward transfer, J. Mater. Res. 26 (2011)2438–2449, http://dx.doi.org/10.1557/jmr.2011.215.

25] C.F. Brasz, C.B. Arnold, H.A. Stone, J.R. Lister, Early-time free-surface flow

driven by a deforming boundary, J. Fluid Mech. 767 (2015) 811–841, http://dx.doi.org/10.1017/jfm.2015.74.

26] M. Duocastella, V. Dinca, J.M. Fernández-Pradas, J.L. Morenza, P. Serra, Studyof liquid deposition during laser printing of liquids, Appl. Surf. Sci. 257 (2011)5255–5258, http://dx.doi.org/10.1016/j.apsusc.2010.10.148.


Recommended