• No results found

3D printed features in the 100 μm range for application in sensing

N/A
N/A
Protected

Academic year: 2021

Share "3D printed features in the 100 μm range for application in sensing"

Copied!
4
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

3D printed features in the 100 µm range

for application in sensing

Jort Verhaar, Remco Sanders and Gijs Krijnen

MESA+ Institute for Nanotechnology

Transducers Science and Technology, University of Twente PO 217 7500 AE Enschede, The Netherlands Email: j.verhaar@student.utwente.nl, gijs.krijnen@utwente.nl

Abstract—In this work the 3D extrusion printing fabrication process for intricate structures is examined. Required support material normally is removed by brute force water jetting. We investigated the chemical dissolution of Fullcure 705 support ma-terial while minimally affecting Fullcure 720 structural mama-terial. From several solvents ethanol turned out to be the best with respect to selectivity and dissolution speed. We found that the dissolution process can be theoretically accurately described by the Noyes-Whitney equation, implying that the development of various structures can be predicted quite well. The fabrication process was used to make various 5 mm diameter membranes, ranging in thickness from 112 to 768 µm and their mechanical performance was characterised.

I. INTRODUCTION

3D printing technology is developing in fast pace, giving possibilities for applications in new fields. One of the new areas of exploration is the fabrication of sensors and actu-ators. Several materials with conductive properties and new fabrication techniques have been developed in order to extend the technology to this field [1], [2]. 3D printing technology has the potential to bridge the gap between MEMS and precision engineered sensors, advantages being the monolithic integration and customisation of the transducer in structural parts, e.g. not only robotic limps and prosthetic devices but any application for which structural parts are already being printed. In order to have mechanical sensors it is advantageous that comparatively small and flexible structures can be printed. In this work we focused on the printing of membranes since these are useful in many applications (microphones, loudspeakers, acoustic windows, mass-sensing, etc) while tech-nologically being challenging at the same time. Our inspiration has been to use membranes to mimic tympanal membranes of the locust ear. In these membranes traveling waves propagate with different speeds depending on frequency and the mechan-ical properties of the membranes, whereas the dimensions and geometry of the membrane determine where the wave will eventually localise, providing a passive mechanical frequency fractionation process [3]–[6]. By using various materials across the membrane the mechanical properties can be tailored. 3D printing technology can provide these different materials and apply them in a range of thicknesses, thus forming in principle an ideal tool for fabrication of such intricate membranes.

Sensors and actuators demand small scale structures. The corresponding maximum thickness of these membranes chal-lenges current print technologies. We printed structures making use of a PolyJet Eden 250 printing Fullcure 720 structures.

x z y

Fig. 1. CAD design of plate containing membranes varying in thickness from 16 to 800 µm.

The models are supported by Fullcure 705 support material. The technique is able to print layers as thin as 16 µm, but the cleaning method needs to be adjusted to rinse structures smaller than 800 µm, which is the minimum wall thickness. Figure 1 shows a CAD-model containing several membranes at various depths.

II. DISSOLUTION OF SUPPORT MATERIAL

In order to examine the possibility to release the support material by chemical dissolution, dissolution tests were carried out. Acetone, ethanol, potassium hydroxide and water where chosen as solvents based on the composition of both print materials. The dissolution process can be described using the Noyes-Whitney equation from [7]–[9].

dm(t)

dt = A(t)D

(Cs Cb(t))

d (1)

The dissolution time rate is limited by the diffusion speed (D) of dissolved material through the boundary layer thickness (d) and the concentration difference. At the material/solvent interface the material dissolves with the maximum saturation concentration (Cs). Dissolved material will diffuse away from

the material to lower concentrations until it reaches a stationary concentration, which is the time dependent bulk concentration (Cb(t)). This bulk concentration is given by the ratio of the

amount of material that was dissolved from the structure into

(2)

0 0.5 1 1.5 2 2.5 3 x 104 0 1000 2000 3000 4000 5000 6000 Time (s) Material loss (mg) Water Acetone Potassium Hydroxide Ethanol

Fig. 2. Material loss for Fullcure 705 support material for Water, Acetone, Potassium Hydroxide and Ethanol.

the solvent and the volume of the solvent. The boundary layer (d) defines the path length needed to reach the bulk concen-tration. When the diffusion constant (D) is multiplied with the area available for dissolution (A) and the concentration gradient, the mass change rate (dm/dt) is found. The diffusion coefficient over the boundary layer thickness is constant for a fixed geometry. This constant is defined as the dissolution rate constant (D/d = k).

To measure the dissolution rate Fullcure 705 boxes of 1 cm3 were dissolved in a bath filled with 200 ml solvent.

Every twenty minutes the material loss of the cubes was measured. To measure the unintentional dissolution of the structural material, the dissolution of Fullcure 720 material was investigated by making use of 2 mm diameter cylinders with a height of 2 cm.

Figure 2 shows the material loss over time for dissolution of Fullcure 705 in several solvents. The fitted solid lines show that the Noyes-Whitney model is a good approximation of the experimental values (circles). Figure 3 shows the mass rate change for Fullcure 720 in the solvents. The maximum saturation concentration Cs and diffusion rate constant k

for the differente material/solvent combinations, as found by fitting the Noyes-Whitney mode, are given in table I.

The figures show that, ethanol hardly effects the Fullcure 720, acetone initially is absorbed by the material and potassium hydroxide slightly dissolves the model material. Indicating that ethanol is the solvent most appropriate for fast cleaning of the model based on its selectivity and dissolution speed.

The dissolution is depending on the shape of the to-be-dissolved material. Shapes used in these comparison tests do not deviate from the Noyes-Whitney model. Objects with a support material area in the printed plane (x y) dissolute layer-wise. This layer-wise dissolution instead of homoge-neous dissolution gives deviations in the predicted time needed for releasing the model.

III. REALIZEDMEMBRANES

The 5 mm diameter membranes consist of multiple 16 µm thick layers. The thinnest membrane produced has a total

0 0.5 1 1.5 2 2.5 3 x 104 400 500 600 700 800 900 1000 Time (s) Material loss (mg) Water Acetone Potassium Hydroxide Ethanol

Fig. 3. Material loss for Fullcure 720 model material for Water, Acetone, Potassium Hydroxide and Ethanol.

TABLE I. DISSOLUTION VALUES OFFULLCURE705ANDFULLCURE 720IN DIFFERENT SOLVENTS. Cs(mg k(cm s 1) Cs(mg k(cm s 1) cm 3) cm 3) Fullcure 705 705 720 720 Water 27.6 424.4⇥ 10 6 0.02 0.0⇥ 10 6 Acetone 181.1 8.696⇥ 10 6 0.04 0.0⇥ 10 6 Ethanol 37.61 1789⇥ 10 6 0.29 0.0⇥ 10 6 Potassium 640.9 4.134⇥ 10 6 0.01 30.5⇥ 10 6 Hydroxide 1 mm x z

Fig. 4. Cross-section of the 112 µm membrane under optical microscope after gold deposition and slicing.

thickness of 112 µm. Other membranes tested have a thickness of 320, 352, 752 and 768 µm. Figure 4 shows a cross-section of the 112 µm membrane captured with the optical microscope. A layer of gold is deposited on the membrane in order to improve reflection under the White Light Interference Microscope (WLIM).

Tests are performed to determine the mechanical perfor-mance and mechanical properties as the Youngs Modulus and intrinsic stress. The ratio between intrinsic stress and bending stiffness defines the mechanical performance according to equation 2 [5].

R =3N

2(1 2)

Et4!2 (2)

(3)

M icr omet er S cr ew w0 x e Scale P

Fig. 5. Measurement setup for bending of membrane under point load.

where E is the Youngs modulus, t the thickness of the membrane, ⇢ the density of the material, N the linear tension, ⌫the Poisson ratio and ! the radial frequency of the vibration. When R ⌧ 1 the bending stiffness dominates and plate behavior is observed. If R 1the linear tension N causes the plate to exhibit membrane behavior. These different behaviors cause their own specific bending profiles (see e.g. figure 8) and mechanical performances.

The relation between a force P applied through a small point at the center of a circular membrane and the maximum deflection w0 is given in equation 3 [10].

P = 4⇡tw0 ✓ T+ 2w2 0E 3R2(1 ⌫) ◆ (3) where T is the intrinsic stress and R is the radius of the

membrane.

A schematic of the measurement setup is shown in figure 5. The membrane is supported by its printed support (red). A micrometer screw is fixed to a bar with a connected 500 µm diameter point. This point is given a displacement by the micrometer screw. A scale under the membrane structure measures the applied force. The error introduced in this measurement due to displacement of the spring in the scale is accounted for in baseline measurements.

In figure 6 the deflection w0is plotted against the applied

force. The corresponding Youngs modulus and intrinsic stress, calculated with equation 3, is shown in table II. The 112 and 320 µm thick membranes are stress dominated, resulting in membrane behaviour. The membranes with 752 µm and 768µm thickness are stiffness dominated resulting in plate behaviour.

The bending of a membrane under static uniform load gives information on the uniformity and characteristics of the realized membrane. In order to investigate the membrane shape under deflection, White Light Interference Microscopy

TABLE II. MEMBRANE PROPERTIES UNDER POINT LOAD. Membrane (µm) E(MPa) T(MPa) Remark

112 1360 10 Membrane behavior 320 1739 16 Membrane behavior 352 48 2 Ruptured membrane 752 660 43 Plate behavior 768 690 16 Plate behavior 0 100 200 300 400 500 600 700 −0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 Force (nN) Deflection w 0 (mm) 112µm 320µm 352µm 752µm 768µm

Fig. 6. Deflection of membranes under point load.

(WLIM) is applied. A layer of gold of 100 nm thickness is applied to the membrane by evaporation in order to increase its optical reflection. In figure 7 the expected bending for a uniform membrane under uniform pressure is shown. The relation between pressure difference and maximum deflection is given in equation 4 [10]. p = 4tw0 R2 ✓ 4 3 Et2 R2(1 2)+ T+ 64 105 Ew2 0 R2(1 2) ◆ (4)

The 112 µm membrane is used for deflection measurements with the WLIM. The membrane is supported in its printed sup-port structure and airtight connected to an air pump. A pressure of 300 mbar is applied to the membrane. The deflection is shown with dots in figure 8. Theoretical deflection of a stress dominated (blue line) and stiffness dominated (green line) are shown. The realised membrane exhibits membrane dominated behaviour. Fitting of the theoretical deflection curve results in a predicted Youngs modulus of 1385 MPa and an intrinsic stress of 19 MPa.

The mechanical performance of the membrane is still ill defined due to deviations in intrinsic stress in the structure. The layer of gold creates a bi layer structure altering the intrinsic stress. Values for the Youngs modulus are equal before and after deposition, whereas the stress increases from 10 to 19 MPa. p α Fs,z a z x

Fig. 7. Bending of membrane under constant pressure

(4)

0.5 1 1.5 2 2.5 3 3.5 0 100 200 300 400 500 600 700 800 x location on membrane (mm) De flection w of membrane (µm) Membrane deflection Plate deflection Measurement Data

Fig. 8. In red dots the measured membrane shape under a uniform load of 300 mbar. The theoretical shapes for membrane (blue) and plate (green) behaviour are compared with the shape of the realised membrane.

IV. DISCUSSION

In literature the Youngs modulus of Fullcure 720 is given to be 2870 MPa, but with variations for larger printed structures between 2000 MPa to 3000 MPa [11], [12]. It appears that the effective Youngs Modulus for our printed structures is much lower. On the other hand the Youngs modulus as determined from both the point loading (1360 MPa) and pressure loading (1385 MPa) measurements are almost identical, giving credi-bility to the measurements.

The difference of the effective Youngs modulus as found in this work and by others may be due to the way in which the 16 µm layers are mechanically connected. From eq. 4 we can see that the relation between deflection and pressure is nonlinear in membrane thickness t. Therefore a membrane consisting of a number of mechanically weakly connected layers will show a different stiffness than a single layer of the same thickness and material. However, the different layers in our samples seem to be coupled, as can be seen in figure 9 left. Moreover, for the pressure loading experiments the deflection in the center w0 is much larger than the total membrane

thickness t causing the first term to become insignificant relative to the third one. Hence, the mechanical connection of the layers constituting the membrane doesn’t affect the determination of the Youngs modulus in this experiment.

Finally, an effect present in our membranes is that the sides to which the membranes are connected seem to be partly undercut in the solution process, see figure 9 right. This leads to a membrane radius which effectively is somewhat larger than designed. Plugging in the determined values for Youngs modulus and intrinsic stress shows that the third term between brackets in eq. 4 is about 5 times larger than the second term and about 20 times larger than the first term under the conditions of the pressure loading experiment. Therefore the pressure difference - displacement relation shows a quartic dependence on radius. Hence an undercut of e.g. 250 µmon a 2.5 mm radius membrane will result in a ⇡ 35 % underestimation of the Youngs modulus. Correcting for this undercut would give a Youngs modulus value of 2110 MPa, much closer to published values [11], [12].

Fig. 9. Left: SEM figure of a cross-section of membrane after slicing. The dotted red lines mark the membrane boundaries. Right: sidewalls of a developed membrane, showing some disconnection of the layers over a short distance from the side.

V. CONCLUSION

3D printing technologies can be used to create membranes at the 100 µm thickness range. For the combination of Fullcure 720/705 materials these membranes can be released from their support material by making use of ethanol dissolution. The process of dissolution can be simulated effectively for the de-signed structures. Ethanol can be used to rinse complex struc-tures in a fast way, acetone can be used for slower dissolution. The membranes posses mechanical properties which can be used beneficially for transducers applications, e.g. to mimic the locust tympanal membrane. Solvents like potassium hydroxide can be used to smoothen surfaces of printed models. For a structure consisting of 7 printed layers the Youngs modulus of the Fullcure 720 structural material ranged between 831 and 1385 MPa. This result can be used to design membranes with travelling wave guiding properties.

REFERENCES

[1] S. J. Leigh, R. J. Bradley, C. P. Purssell, D. R. Billson, and D. A. Hutchins, “A simple, low-cost conductive composite material for 3d printing of electronic sensors,” PLoS ONE, vol. 7, no. 11, p. e49365, 11 2012.

[2] E. MacDonald, R. Salas, D. Espalin, M. Perez, E. Aguilera, D. Muse, and R. Wicker, “3d printing for the rapid prototyping of structural electronics,” Access, IEEE, vol. 2, pp. 234–242, Dec 2014.

[3] A. Michelsen, “The physiology of the locust ear,” Zeitschrift f¨ur vergleichende Physiologie, vol. 71, no. 1, pp. 102–128, 1971. [4] R. Malkin, T. R. McDonagh, N. Mhatre, T. S. Scott, and D. Robert,

“Energy localization and frequency analysis in the locust ear,” Journal of The Royal Society Interface, vol. 11, no. 90, 2013.

[5] P. Westerik, “Mems acoustic sensors inspired by the locust ear,” Master’s thesis, University of Twente, 2014.

[6] P. Westerik, E. Berenschot, and G. Krijnen, “Development of a biomimetic eardrum for acoustic sensing,” in IEEE Sensors. IEEE, November 2015.

[7] J. Siepmann and F. Siepmann, “Mathematical modeling of drug disso-lution,” International Journal of Pharmaceutics, vol. 453, no. 1, pp. 12 – 24, 2013, poorly Soluble Drugs.

[8] A. Noyes and W. Whitney, “The rate of solution of solid substances in their own solutions,” J. Am. Chem. Soc, vol. 19:930-934, 1897. [9] H. Mehrer, Diffusion in Solids: Fundamentals, Methods, Materials,

Diffusion-Controlled Processes, ser. Springer Series in Solid-State Sci-ences. Springer, 2007.

[10] W. Schomburg, Introduction to Microsystem Design. Springer-Verlag Berlin Heidelberg, 2011.

[11] Objet Geometries Ltd., “Fullcure materials,” 2009.

[12] Stratasys Ltd., “PolyjetTMmaterials data sheet,” stratasys.com website, visited 2015.

Referenties

GERELATEERDE DOCUMENTEN

It is shown that the sensor is able to measure resistance changes due to deflections, but the wind tunnel tests did not show fluctuating behaviour or significant changes in

The chemical compositions of the support material and model material are used to determine a suitable solvent for dissolving the support material.. Fullcure 705 and fullcure 720

Explain the principles of control planes in optical networks: static versus dynamic, dynamic overlay versus dynamic peer.. Explain: FDMA, TDMA, SDMA

Van Wageningen and Du Plessis (2007), analysing 5-min rainfall data for the Molteno reservoir rainfall station in Cape Town in the Western Cape over the period 1961–2003, found

Hoewel het lithisch materiaal niet in situ werd aangetroffen, kunnen er nog sporen of vondstlagen uit deze perioden bewaard zijn gebleven.. Ook de talrijke aanwezige bodemsporen

We found that the promoter regions for 2343 out of 6270 yeast genes (37%) are bound by one or more of the 106 transcriptional regulators in yeast cells grown in rich medium when

Since several parts of a language implementation depend on the intermediate language, this approach requires the implementation of a dedicated tool chain: A compiler map- ping