Skip to content
BY-NC-ND 4.0 license Open Access Published by De Gruyter March 8, 2017

PDMS electrodes for recording and stimulation

  • Alexander Brensing EMAIL logo , Roman Ruff , Benjamin Fischer , Sascha L. Wien and Klaus-Peter Hoffmann

Abstract:

The usability of flexible electrodes in moving environment is limited due to different mechanical characteristics of their metallic and polymeric components. To achieve structure compatible electrodes, all used materials need to have similar Young’s moduli as the surrounding tissue. This paper describes the characterization of macroscopic as well as miniaturized electrodes entirely made out of modified silicone (PDMS). Electrochemical, mechanical, biological, optical, and applicative methods were used. It could be shown, that PDMS electrodes are capable to be used for recording electrocardiograms with similar form and amplitude as with standard electrodes.

1 Introduction

Electrodes serve as electrical interface between biological tissues and technical systems. They are capable for recording bioelectrical potentials or stimulation of tissue. Surface electrodes as well as implantable ones are used in many sectors of medicine and medical engineering. Nevertheless, an application of flexible electrodes in moving muscular environment causes mechanical stress within the structures and to the tissue. Cardiac muscles changes their length up to about 10% [1], other types even more [2]. The Young’s moduli (YM) of the tissue and the materials used for flexible electrodes often distinguish from each other. The conductive structures of electrodes are usually made out of metals (e.g. gold, YM 78 GPa [3]) or intrinsic conductive polymers (e.g. PEDOT:PSS, YM 1 GPa [4], [5]). The use of polymeric substrates like polyimide (PI, YM 10 GPa [6]) or polydimethylsiloxan (PDMS, YM 9 MPa [7]) provides the flexibility [5], [8] . The Young’s modulus of muscle tissue is in the range of 10−3 Pa (not active) to 20 MPa (tetanic) [9]. These differences limit the long-term stability of micro electrodes used in moving tissue. Hence, the development of stretchable and thus structure compatible electrodes is of high interest.

This work was aimed to characterize electrodes entirely made out of PDMS as monomorphic material as well as the used raw materials.

2 Methods

The conducting paths of the electrode were made of a PDMS composite (MED-6015, Nusil) including carbon nanotubes (CNT, Sigma-Aldrich). Using them to form the composite material, arose in an electrical resistance, that was independent of elongation [10]. Different amounts of CNT were compared regarding the specific conductivity of the resulting composite. For the electrode contact, an established material for dry surface electrodes was used [11]. To reduce the YM of the materials, a silicone thinner (ST, Smooth-ON) was added. A soft lithographic fabrication process was used to connect all PDMS composites as different components of the electrode. For structuring the conduction paths, a screen printing based technique was used. Detailed information for this method can be found elsewhere [12]. Two electrode designs – a macroscopic monopolar and a miniaturized bipolar version – were investigated (see Figure 1).

Figure 1 PDMS electrodes (top: macroscopic electrode, bottom: miniaturized electrode, also shown in [12]).
Figure 1

PDMS electrodes (top: macroscopic electrode, bottom: miniaturized electrode, also shown in [12]).

To characterize the electrodes and base materials electrical, electrochemical, mechanical, optical, biological and applicative methods were used. The specific conductivity of the CNT-PDMS material was measured by using a four point setup. For mechanical characterization, i.e. to determine the YM, tensile tests were carried out. The homogeneity of the particles within the polymer matrix was visualized with confocal Raman microscopy [13]. To evaluate the technical characteristics of the electrodes, electrochemical impedance spectroscopy, cyclic voltammetry (CV), and pulse tests (PT) were performed. Therefor, a three electrode setup in 0.9% sodium chloride solution was used. Using CV the voltage boundaries (VB) of both macroscopic and miniaturized electrodes were determined. Based on these values, the charge injection capacity (CIC) could be evaluated by analysis of the voltage response to current pulses (details in [14] and [15]). The current was measured to calculate the current density at the electrode surface. It was determined by summation of the discrete values of the current pulses and dividing it by the sampling frequency (2.5 Hz). The tolerance to cells could be shown by direct cell contact tests. Therefor, human lung fibroblasts (LGC Standards) were used. Additionally the electrodes were colonized with human induced pluripotent stem cell derived cardiomyocytes (hIPS-CM, AxioGenesis) to track their behavior over time. This required a preliminary coating of the electrodes with matrigel (BD Biosciences). Electrocardiograms (ECG) were recorded over extremity acquisition (Einthoven I method). For comparison Ag/AgCl-electrodes were used.

3 Results

The specific conductivities of the CNT-PDMS composites are shown in Figure 2. Due to the addition of CNT into the PDMS a specific conductivity of (31.5 ± 0.9) S/m could be achieved. In wide ranges the CNT were homogeneously dispersed, but in the inner of the material little aggregations of silicone could be observed (see Figure 3). The YM of the different materials are listed in Table 1.

Figure 2 Specific conductivity of CNT-PDMS at different particle concentrations in the range of good processability.
Figure 2

Specific conductivity of CNT-PDMS at different particle concentrations in the range of good processability.

Figure 3 Raman microscopic images (left, red: CNT signal, green: PDMS signal; scale bar: 50 µm) and correlations graph (right) of a cross section at the border of the conductor material.
Figure 3

Raman microscopic images (left, red: CNT signal, green: PDMS signal; scale bar: 50 µm) and correlations graph (right) of a cross section at the border of the conductor material.

Table 1

Young’s moduli (YM) of the electrode materials and standard deviation; also presented in [12].

MaterialYM in MPa
Substrate (N = 6)0.99 ± 0.02
Conductor (N = 13)1.00 ± 0.10
Electrode contact (N = 9)4.57 ± 0.17

The impedance spectra associated to the electrodes are shown in Figure 4. For frequencies between 10 Hz and 25 kHz, an ohmic (i.e. frequency independent) behavior was observed for both electrode sizes. The impedance magnitude of the macroscopic electrodes (diameter 1 cm) lay between 3 kΩ and 35 kΩ (median 5 kΩ). The electrodes with a diameter of 0.4 cm had impedances in the range between 6 kΩ and 102 kΩ (median 37 kΩ). The CV did not provide a clear water window as for metallic electrodes. Hence, as VB values were chosen, at which the CV did not possess peaks, and thus no signs for chemical reactions were apparent. The negative boundary was defined as the particular voltage turning point, at which in case it was overrun, the first overlap of the curves occurred (see Figure 5, left). The positive potential turning point whereby exceeding a maximum arises within the curve was defined as positive VB (see Figure 5, right). Potentials larger than the VB may firstly result in reversible processes. Visible hydrolysis could not be observed, even at potentials up to ±3 V. The PT on exemplary demonstrator electrodes yielded to electrode potentials shown in Figure 6. The resulting VB and CIC are summarized in Table 2.

Figure 4 Impedance spectra (top: macroscopic electrode, bottom: miniaturized electrode).
Figure 4

Impedance spectra (top: macroscopic electrode, bottom: miniaturized electrode).

Figure 5 Cyclic voltammograms for evaluation of the voltage boundaries (VB) of a macroscopic electrode; left: negative VB (−0.5 V), right: positive VB (1.1 V); effects occurring at overrun are marked by arrows.
Figure 5

Cyclic voltammograms for evaluation of the voltage boundaries (VB) of a macroscopic electrode; left: negative VB (−0.5 V), right: positive VB (1.1 V); effects occurring at overrun are marked by arrows.

Figure 6 Minimal and maximal electrode potentials measured as voltage response to a biphasic current impulse vs. the related current densities at the electrode surface; left: macroscopic electrode, right: miniaturized electrode.
Figure 6

Minimal and maximal electrode potentials measured as voltage response to a biphasic current impulse vs. the related current densities at the electrode surface; left: macroscopic electrode, right: miniaturized electrode.

Table 2

Voltage boundaries (VB) and charge injection capacities (CIC, ± standard deviation; also in [12]) of macroscopic (diameter 1 cm) an miniaturized (diameter 0.4 cm) electrodes.

ElectrodeVB in VCIC in μC/cm3
Macroscopic−0.5 to 1.10.13 ± 0.02
Miniaturized−0.6 to 1.31.40 ± 0.34

In direct cell contact non of the used materials caused any visible cell damage or alteration. Furthermore, hIPS-CM attached successfully to the electrode surface. In Figure 7, a dense cell layer can be seen. Videos of beating cells were recorded and analyzed over 24 days. The cells grown on the electrodes regularly beat with frequencies between 2.8 ± 0.2 bpm and 18.9 ± 4.7 bpm. After 10 days no beating could be observed any more. At day 13 and day 24 live/dead staining of cells cultured on the electrodes were performed. For comparison all investigations were also performed with hIPS-CM cultured on conventional polystyrene substrates. The mean beating frequency of this cell control was 2 times higher than that of the cells at the electrodes. After 13 days approximately half of the cells were still alive. At day 24 all cells on the electrodes were dead, while the cells of the control entirely were alive. The physiological ECG, shown in Figure 8, visualize the comparability of the PDMS electrodes to standard Ag/AgCl electrodes. The signals of all three electrode types had a similar form and amplitude.

Figure 7 Transmission microscopic view of hIPS-CM grown directly next to a PDMS electrode (day 7, electrode contact in the lower right corner).
Figure 7

Transmission microscopic view of hIPS-CM grown directly next to a PDMS electrode (day 7, electrode contact in the lower right corner).

Figure 8 ECG (Einthoven I) comparison between macroscopic and miniaturized electrodes (PDMSmacro and PDMSmini) and standard electrodes (Ag/AgCl).
Figure 8

ECG (Einthoven I) comparison between macroscopic and miniaturized electrodes (PDMSmacro and PDMSmini) and standard electrodes (Ag/AgCl).

4 Conclusion

Macroscopic and miniaturized PDMS electrodes were characterized. Electrically conductive and non-conductive silicone composites were used as materials for all electrode components. The YM of the polymers and (tetanic) muscle tissue were in the same range. The cell compatibility of the components to lung fibroblasts could be shown. Also, high sensitive hIPS-CM retained their vitality for more than 10 days in direct contact to the electrodes. Hence, for the electrodes a structure and surface compatibility has been achieved. The beating frequency of hIPS-CM has been in the same range as published in literature [16], but also lower beat rates were observed. This may be due to an non optimal distribution or absorption on the matrigel layer. Compared to metal electrodes of similar size (area 1 cm2 [15]) the PDMS electrodes (area 0.79 cm2) show a much higher impedance for frequencies larger than 1 Hz (three orders of magnitude at 10 kHz). Although the median values of the macroscopic and the miniaturized electrodes were 5 kΩ and 37 kΩ, respectively, the acquired ECG signals had a similar form and amplitude compared to the recordings of standard electrodes. The determined CIC values are much lower than that of metallic micro electrodes (Pt electrodes, diameter 300 μm, CIC 64 μC/cm2 [15]). However, 2008 Cogan described in his review the CIC as indirect proportional to the size electrode [17].

It could be shown, that PDMS electrodes can be used for recording physiological signals and for stimulation purposes. The fabrication process [12] is capable for production of functional miniaturized and macroscopic electrodes, both as bipolar and monopolar structures.

Author’s Statement

Research funding: This work was supported by the project elaN “Micro Nano integration as key technology for the next generation of sensors and actuators” (16SV5367), funded by the German Federal Ministry for Education and Research (BMBF). Conflict of interest: Authors state no conflict of interest. Informed consent: Informed consent has been obtained from all individuals included in this study. Ethical approval: The research related to human use complies with all the relevant national regulations, institutional policies and was performed in accordance with the tenets of the Helsinki Declaration, and has been approved by the authors’ institutional review board or equivalent committee.

References

Ter Keurs HE, Shinozaki T, Zhang YM, Zhang ML, Wakayama Y, Sugai Y, et al. Sarcomere mechanics in uniform and non-uniform cardiac muscle: A link between pump function and arrhythmias. Prog Biophys Mol Biol. 2008;97:312–31.10.1016/j.pbiomolbio.2008.02.013Search in Google Scholar PubMed

Arnold EM, Delp SL. Fibre operating lengths of human lower limb muscles during walking. Philos Trans R Soc B Biol Sci. 2011;366:1530–9.10.1098/rstb.2010.0345Search in Google Scholar PubMed PubMed Central

Pamula VK, Jog A, Fair RB. Mechanical property measurement of thin-film gold using thermally actuated bimetallic cantilever beams. In: Technical Proceedings of the 2001 International Conference on Modeling and Simulation of Microsystems. Hilton Head Island; 2001.Search in Google Scholar

Greco F, Zucca A, Taccola S, Menciassi A, Fujie T, Haniuda H, et al. Ultra-thin conductive free-standing PEDOT/PSS nanofilms. Soft Matter. 2011;7:10642–50.10.1039/c1sm06174gSearch in Google Scholar

Green RA, Matteucci PB, Hassarati RT, Giraud B, Dodds CW, Chen S, et al. Performance of conducting polymer electrodes for stimulating neuroprosthetics. J Neural Eng. 2013;10:016009.10.1088/1741-2560/10/1/016009Search in Google Scholar PubMed

Mittal KL. Polyimides and other high temperature polymers: synthesis, characterization and applications. Volume 2. Utrecht: VSP; 2003.10.1201/9789047403302Search in Google Scholar

Schmid H, Michel B. Siloxane polymers for high-resolution, high-accuracy soft lithography. Macromolecules 2000;33:3042–9.10.1021/ma982034lSearch in Google Scholar

Stieglitz T. Considerations on surface and structural biocompatibility as prerequisite for long-term stability of neural prostheses. J Nanosci Nanotechnol. 2004;4:496–503.10.1166/jnn.2004.075Search in Google Scholar PubMed

Lohmann S. Eigenschaften biologischer Materialien zur Simulation menschlicher Bewegung – Funktionell-anatomische Grundlagen und Materialeigenschaften zur Erstellung dreidimensionaler Körpermodelle im Computer. Dissertation. Konstanz: Konstanz University; 2005.Search in Google Scholar

Brensing A, Hauck S, Ruff R, Poppendieck W, Hoffmann K-P. Comparison of different conductive fillers in silicone for the purpose of replacing metallic conductive structures in flexible implants. Biomed Tech. 2013;58/SI-1/Track-O.10.1515/bmt-2013-4382Search in Google Scholar PubMed

Hoffmann KP, Ruff R. Flexible dry surface-electrodes for ECG long-term monitoring. In: 29th Annual International Conference of the IEEE. Eng Med Biol Soc; 2007. p. 5739–42.10.1109/IEMBS.2007.4353650Search in Google Scholar PubMed

Brensing A, Ruff R, Hoffmann KP. Manufacture of electrodes entirely made of modified PDMS. J Sens Sens Syst. 2017;6.10.5162/sensor2017/P1.1Search in Google Scholar

Kreiner-Møller A, Stracke F, Zimmermann H. Confocal Raman microscopy as a non-invasive tool to investigate the phase composition of frozen complex cryopreservation media. Cryo-Letters. 2013:34:248–54.Search in Google Scholar

Rose TL, Robblee LS. Electrical stimulation with Pt electrodes. VIII. Electrochemically safe charge injection limits with 0.2 ms pulses (neuronal application). IEEE Trans Biomed Eng. 1990;37:1118–20.10.1109/10.61038Search in Google Scholar PubMed

Poppendieck W. Untersuchungen zum Einsatz neuer Elektrodenmaterialien und deren Evaluation als Reiz- und Ableitelektrode. Dissertation. Saarbrücken: Saarland University; 2009.Search in Google Scholar

Rajasingh S, Thangavel J, Czirok A, Samanta S, Roby KF, Dawn B, et al. Generation of functional cardiomyocytes from efficiently generated human iPSCs and a novel method of measuring contractility. PLoS One 2015;10:e0134093.10.1371/journal.pone.0134093Search in Google Scholar PubMed PubMed Central

Cogan SF. Neural stimulation and recording electrodes. Annu Rev Biomed Eng. 2008;10:275–309.10.1146/annurev.bioeng.10.061807.160518Search in Google Scholar PubMed

Published Online: 2017-3-8
Published in Print: 2017-3-1

©2017 Alexander Brensing et al., licensee De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/cdbme-2017-0014/html
Scroll to top button