Brought to you by:
Paper

Characterization of spiral ganglion neurons cultured on silicon micro-pillar substrates for new auditory neuro-electronic interfaces

, , and

Published 28 January 2015 © 2015 IOP Publishing Ltd
, , Citation M Mattotti et al 2015 J. Neural Eng. 12 026001 DOI 10.1088/1741-2560/12/2/026001

1741-2552/12/2/026001

Abstract

Objective. One of the strategies to improve cochlear implant technology is to increase the number of electrodes in the neuro-electronic interface. The objective was to characterize in vitro cultures of spiral ganglion neurons (SGN) cultured on surfaces of novel silicon micro-pillar substrates (MPS). Approach. SGN from P5 rat pups were cultured on MPS with different micro-pillar widths (1–5.6 μm) and spacings (0.6–15 μm) and were compared with control SGN cultures on glass coverslips by immunocytochemistry and scanning electron microscopy (SEM). Main results. Overall, MPS support SGN growth equally well as the control glass surfaces. Micro-pillars of a particular size-range (1.2–2.4 μm) were optimal in promoting SGN presence, neurite growth and alignment. On this specific micro-pillar size, more SGN were present, and neurites were longer and more aligned. SEM pictures highlight how cells on micro-pillars with smaller spacings grow directly on top of pillars, while at wider spacings (from 3.2 to 15 μm) they grow on the bottom of the surface, losing contact guidance. Further, we found that MPS encourage more monopolar and bipolar SGN morphologies compared to the control condition. Finally, MPS induce longest neurite growth with minimal interaction of S100+ glial cells. Significance. These results indicate that silicon micro-pillar substrates create a permissive environment for the growth of primary auditory neurons promoting neurite sprouting and are a promising technology for future high-density three-dimensional CMOS-based auditory neuro-electronic interfaces.

Export citation and abstract BibTeX RIS

Abbreviations

DIV days in vitro
MPS micro-pillar substrates
SGN spiral ganglion neurons
CI cochlear implant

1. Introduction

Cochlear implants (CIs) aim to partially restore the sensation of sound in deaf people. CIs are implanted in people with damaged hair cells, where an electrode array is used to stimulate the auditory nerve. This system consists of an external part that is responsible for the capture, encoding and transmission of sound information to the internal, surgically implanted electrode array (Waltzman and Cohen 2000). CIs often result in good outcomes, with high speech perception scores and increased joy of listening to music, albeit typically only in ideal conditions with minimal background noise. However, the efficacy of the electrical stimulation of speech via a CI is highly variable, with speech perception scores in standard audiological assessments ranging between zero and almost perfect performance (Wilson and Dorman 2008a, 2008b, Blamey et al 2001). Such enormous variability is observed even in simple perceptual tasks such as voice gender identification (Kovacic and Balaban 2009).

In the human cochlea, approximately 3400 inner hair cells, playing the role of auditory sensory receptors, innervate about 35 000 spiral ganglion neurons (SGN), which are subsequently connected to the neurons of the cochlear nucleus in the brain stem (Liu et al 2012, Spoendlin and Schrott 1990, Rask-Andersen et al 2012). CIs are mainly based on a tonotopic organization of the auditory system, establishing a direct frequency-place map between the acoustical frequency of the stimulating sound and the position of the maximally stimulated cells in the cochlea, which is maintained throughout the whole auditory pathway (Bear et al 2006, Loizou et al 1998, Greenwood 1990). A CI bypasses damaged inner hair cells and thus SGN survival is crucial for the CI to be an effective neuro-prosthetic device. Mimicking naturally existing structures, miniaturizing the electrode size and increasing the electrode number is one promising but challenging approach to improving the neuro-electronic interface to enhance the efficacy of CIs in a real communication settings (Wilson and Dorman 2008a, 2008b).

Several factors limit the auditory performance in CI users. One major factor is the hearing history of a CI user that determines the biological aspects of hearing, such as SGN neuronal survival (Kovacic & Balaban 2010, Shepherd & Hardie 2001). Another important factor is the technological aspects of CIs that determine how well the sound is transmitted and encoded to the auditory nerve. For example, the number and size of electrode contacts positioned along the electrode array is considered to be one of the major determinants of auditory performance (Finley et al 2008, Friesen et al 2001). Currently, electrode arrays are made of a platinum-iridium (Pt-Ir) alloy and may consist of up to 22 electrode contacts per CI. It is believed that each of the electrodes stimulates approximately a few hundred or even thousands of auditory nerve fibers (Wilson and Dorman 2008a, 2008b). Moreover, cross-channel interactions arising from overlaps in stimulation patterns of adjacent electrode contacts can lead to non-focal and non-tonotopic stimulation. This significantly decreases the spatial stimulation resolution and thus limits the transmission of fine spectro-temporal cues. Strategies to improve the neuro-electronic interfaces, based on recent advances in tissue engineering and nanotechnology are focused on significantly increasing the number of properly sized electrodes in order to establish an intimate contact between the electrodes and the neural tissue (O'Leary et al 2009). Currently, neural engineering approaches allow the use of silicon micro-fabrication technology. For example, silicon microelectrodes have been tested recently for compatibility with cochlear nucleus neuronal growth (Mlynski et al 2007, Clarke et al 2011, Rak et al 2011, Tuft et al 2013) and have been successfully employed for retinal prosthesis (Kim et al 2009).

The introduction of micro-patterns on substrates has been shown to be beneficial for axonal growth and guidance in vitro. Micro-patterns are known to positively influence neuronal attachment, axonal development and guidance. In some cases micro-patterns can also have an influence on the cell differentiation state, making it more permissive for regeneration (Schmidt and Leach 2003, Hoffman-Kim et al 2010, Clarke et al 2011). In two recent publications the effect of micro-patterned surfaces on SGN and glial cells has been studied. Clarke et al and Tuft et al use methacrylate with micro-grooves. In both studies it is shown that micro-patterns induce high neuronal alignment (Clarke et al 2011, Tuft et al 2013). It is clear that the integration of topographical cues in a silicon substrate may promote growth and guidance of auditory neurons. Refining electrode manufacturing using silicon micro-pillar electrode arrays could be an approach to improve the neuro-electronic interface and to promote neuronal contact and integration with the electrode array.

In this study we tested the performance of silicon micro-pillar substrates (MPSs) as directly related to microelectrode arrays (MEAs) toward the development of a new applied technology for cochlear implants. We used in vitro cultures of SGN from rat pups as a biological model to assess the effect of the material and the micro-pattern on neuronal presence, morphology and alignment.

2. Materials and methods

2.1. Micro-pillar substrates (MPSs) fabrication

Micro-pillars were introduced in silicon substrates as described previously by Micholt et al (Micholt et al 2013). Briefly, in a cleanroom (Class 1000) environment, standard 8 inch silicon wafers were covered by a low temperature oxide layer, followed by a sintering step at 455 °C. Areas with diameters down to 1 μm and a minimal spacing of 600 nm were defined using standard I-line lithography. 3 μm high pillars were created by a timed reactive ion etch step (RIE), followed by 'piranha' (1:3 H20:H2S04) cleaning. The substrate consisted of individual areas micro-patterned with hexagonal pillars of different dimensions. The pillar widths ranged from 1–5.6 μm (1, 1.2, 1.4, 1.6, 1.8, 2, 2.4, 2.8, 4, 5.6 μm) in the vertical direction, while the spacing ranged from 0.6–15 μm (0.6, 0.8, 1, 1.2, 1.4, 1.6, 1.8, 2.0, 2.4, 3.2, 4, 5, 7, 10, 15 μm) as shown in figure 1(A). Separate substrates were then diced in samples of 8 × 8 mm2 to be used for cell culturing.

Figure 1.

Figure 1. Effect of pillar spacing on neuronal presence. (A) The sketch of the MPS layout shown from the top. The MPS design consisted of 150 areas of vertically protruding silicon oxide pillars with different widths and spacings. The pillar width ranged from 1–5.6 μm (1, 1.2, 1.4, 1.6, 1.8, 2, 2.4, 2.8, 4, 5.6 μm) in the vertical direction, while the spacing ranged from 0.6–15 μm (0.6, 0.8, 1, 1.2, 1.4, 1.6, 1.8, 2.0, 2.4, 3.2, 4, 5, 7, 10, 15 μm) in the horizontal direction. The pillar height was kept constant at 3 μm throughout all MPS samples. (B). Scanning electron microscopy images of three representative MPS (Micholt et al 2013). The pillar dimensions are indicated on top of each panel where W refers to the pillar width, while S refers to the pillar spacing. Scale bars are 10 μm. (C). Surface plot of SGN presence on MPS as a function of pillar width and spacing. SGN presence is expressed as cumulative percentages of the total number of 906 neurons obtained from three independent experiments yielding 12 samples in total (see section 2.6). (D). SGN presence as a function of pillar spacing. The percentage of SGN in each bin was calculated for each experiment and averaged (N = 3). The whisker represents 1 SD. ** = p < 0.01 to all other conditions (one-way ANOVA).

Standard image High-resolution image

2.2. Spiral ganglion neurons (SGN) isolation

The institutional animal care and use committee at the University of Split approved all protocols and experimental plans used in this study. Spiral ganglia dissection and culturing protocols were adapted from (Szabo et al 2003, Bostrom et al 2007, Richardson et al 2007, Vieira et al 2007, Clarke et al 2011). Spiral ganglion neurons (SGN) were isolated from postnatal day 5 rats (P5) and decapitated under cold anaesthesia after placing them on ice. Phosphate buffer saline (PBS) supplemented with 0.3% bovine serum albumine (BSA, Sigma-Aldrich) and 0.6% glucose (Sigma-Aldrich) was used as the dissecting buffer (Mattotti et al 2012). The skull was opened mid-sagitally under an operating microscope and the brain was removed. The temporal bone was harvested and transferred to clean dissecting buffer. Then the otic capsule was dissected, and the cochlea were identified and isolated. The organ of Corti and modiolar cartilage were removed and the spiral ganglia collected in the dissecting buffer.

2.3. Preparation of micro-pillar substrates (MPSs) and glass coverslips

Silicon micro-pillar substrates (MPSs) were cleaned overnight with acetone (Merck), sprayed with 70% ethanol (Medimon) and left to dry under sterile conditions. Glass coverslips were sterilized by autoclaving at 120 °C for 20 min. All samples were then placed in 24 well plates (TPP) for testing. All substrates were coated with 0.01% poly-D-ornithine (WT 30 000–70 000, Sigma) at room temperature overnight, cleaned with sterile water and allowed to dry.

2.4. Dissociated SGN culture

The dissociation of SGN was performed enzymatically in 0.25% trypsin-EDTA (Sigma-Aldrich) at 37 °C for 20 min. The trypsinization was stopped by adding an equal volume of DMEM:F12 (Gibco) supplemented with 10% of Fetal Bovine Serum (FBS, Biochem). DNase (Sigma) was also added at 38 U ml−1. The tissue was then triturated; large pieces were allowed to settle down and the cell suspension was collected. Trypsinization was repeated one more time for non-triturated tissue pieces. The cell suspension was then centrifuged for 5 min at 1000 rpm and the pellet was resuspended in the culture medium, which consisted of Neurobasal-A (Gibco) containing 1% Pen-Strep (Lonza), 0.5 mM L-Glutammine (Gibco), B27 (Gibco) supplement and 30 ng ml−1 GDNF (rHu GDNF, AppliChem GmbH). In order to avoid potential interactions among multiple neurotrophic factors, as well as to keep the experimental design simple, the cultures were supplemented by a single neurotrophic factor. We chose GDNF as the preferred neurotrophic factor as it was found to strongly promote neurite outgrowth in P5 rat pup spiral ganglion explants (Euteneuer et al 2013) and to be the most effective among other neurotrophic factors in adult SGN (Boström et al 2010). Cells were counted and seeded in 100 μl volume at a density of 20*103 cells/sample (≈300 cells mm−2). Cells were allowed to settle for 10–15' in the incubator, then the rest of the medium was added carefully in the well. Half of the medium was changed every 2–3 days.

2.5. Immunocytochemistry and cell imaging

SGN were cultured during 1 and 7 DIV and fixed with 4% paraformaldehyde for 30 min. For immunocytochemical analysis, samples were washed three times with PBS, permeabilized with 0.1% Triton-X (Calbiochem) for 5 min and blocked with PBS containing 1% goat serum (GS, Biochem) for 90 min. The cells were then incubated at room temperature for 90 min in PBS 1% GS with mouse monoclonal anti-βIII tubuline (1:200, Millipore) to stain neurons (named Tuj1+ cells afterwards), and rabbit polyclonal anti-S100 (1:200, Sigma-Aldrich) to stain glial cells (named S100+ cells). Cells were then washed three times with PBS, incubated for 90 min in PBS 1% GS at room temperature with Alexa 488 goat anti-mouse (1:500, Molecular Probes), Texas Red goat anti-rabbit (1:500, Santa Cruz) secondary antibodies and 5 μg ml−1 DAPI (Sigma-Aldrich). The samples were finally washed three more times with PBS and prepared for imaging with Imumount (Thermo Scientific). Imaging was carried out under an Olympus DP71 fluorescent microscope equipped with an Olympus U-RFL-T burner.

2.6. Assessment of SGN presence on MPS

We first determined SGN presence on MPS depending on pillar dimension. Surface plots of SGN distribution on MPSs as function of pillar width and spacing were generated with SigmaPlot software with the aim of considering the cumulative percentages. Cumulative percentages were obtained from the cumulative frequency distribution (sum of SGN in each pillar area among different samples), divided by the total number of neurons and multiplied by 100. Subsequently, we grouped the pillar features according to spacing (S) into 4 bins (0.6–1.0 μm; 1.2–2.4 μm; 2.4–5.0 μm and 7.0–15.0 μm). The percentage of SGN in each bin was calculated for each experiment and averaged. Then we assessed the SGN presence on MPSs and compared it to the control glass coverslips. To quantify the total neuronal cell number, Tuj1+ cells were counted under the microscope on each sample (MPSs or glass coverslips) and averaged. Values were then normalized by sample surface area, which was 64 mm2 for MPS and 132 mm2 for control glass.

2.7. Neuronal morphometric analysis

To describe the effect of the dimension of MPS on SGN morphology, morphometric analysis was carried out with the use of ImageJ software (NIH). A minimum of eight pictures for each sample was considered for analysis. We focused on three parameters: longest neurite length, number of sprouting, and neurite alignment.

2.8. Longest neurite length

Among the growing processes of an SGN, we focused on analyzing the longest neurite of each cell as morphologically most resembling an axon. However, we maintained this nomenclature in the text to avoid confusion. The longest neurite length was described in function of time (1DIV, 7DIV) and pillar spacing on MPS. Glass coverslips were considered as the control. The measurements of the longest neurite length were obtained by overlaying a segmented line on images of Tuj1+ SGN and quantifyied with the measuring tool on the calibrated images. The values were then represented in a Box-and-Whisker plot, where bars represent the minimum and the maximum values of the data distribution, the boxes are the interquartile range and the central lines indicate the median values.

2.9. Number of sproutings and neuronal morphology

The number of sproutings and the relation to neuronal morphology was determined on MPSs and control samples as a function of time (1DIV, 7DIV). To determine whether neurons were neurite-free, mono-, bi- or multi-polar, the number of neurites was scored manually in pictures for each Tuj1+ cell, following a classification described previously (Vieira et al 2007, Whitlon et al 2006). The values from the pictures of a single sample were summed together and the percentage calculated with respect to the total number of neurons. Percentages from different samples from the same conditions were averaged and plotted.

2.10. Neuronal alignment

SGN alignment on MPSs as a function of pillar spacing was determined using the ImageJ Fast Fourier Transform (FFT)-Oval Profile plugin as described previously. Briefly, gray-scale images of Tuj1+ neurons were aligned to the horizontal bottom line and subsequently processed with the FFT function of the ImageJ software. The intensity and distribution of the pixels of the resulting image is related to the directional content. The 'Oval profile' plug-in performed the sum of the pixel intensities along a straight line from the centre to the edge of the FFT image (at angle θ, 0° ≤ θ ≤ 180°), which quantifies the relative contribution of objects oriented in that direction. Plots of the sum of pixel intensities along θ represent the directionality of the original image. Constant pixel intensities, independent of direction, represent random oriented images, while peaks along one specific direction represent an image preferentially aligned in that direction (Alexander et al 2006, Mattotti et al 2012). The values from the analyses of the orientation of each neuron were averaged between neurons in the same bin and plotted in the figure.

2.11. SGN neurite interaction with glial cells

Finally, we assessed the potential interaction of Tuj1+ with glial S100+ cells as manifested by being directly overlapped, both on MPSs and the control samples, as a function of time (1DIV, 7DIV). We defined two categories: SGN having the longest neurite in touch with S100+ glial cells and SGN not having an axon in touch with S100+ glial cells. The proximity of Tuj1+ cells with S100+ cells suggests an interaction between the two cell types. The S100+/Tuj1+ ratio and the number of axons of Tuj1+ cells in touch with S100+ glial cells were counted manually in fluorescent, double stained images.

2.12. Scanning electron microscopy (SEM)

For SEM imaging, samples were fixed in 4% paraformaldehyde for 30 min, washed three times, stained with osmium tetroxide for at least 2 h. Subsequently they were immersed twice over the course of 10 min in increasing concentrations of ethanol (steps of 10–30–50–70–90–100% EtOH) and brought to critical point dehydration. Then they were observed using a Philips XL30 scanning electron microscope.

2.13. Statistical analysis

For statistical analysis, data were transferred to spreadsheets and analyzed with the Statgraphics software (StatPoint Technologies Inc. Warrenton, US). Each experiment was repeated at least three times (see figure captions for details). Values in the plots were reported as averages with standard deviation bars, with an exception for the neurite length analysis (figure 4). To test the variance between multiple samples, one-way ANOVA was performed, followed by a post-hoc LSD test. A p value less than 0.05 was considered significant.

3. Results

3.1. Micro-pillars with different geometries influence SGN presence

We assessed growth, behaviour and differentiation of spiral ganglion neurons (SGN) cultured on silicon micro-patterned substrates (MPSs). These silicon surfaces consisted of 150 different areas of 3 μm high micro-pillars with different widths and spacings. The shape of the pillars and their organization was designed in a hexagonal manner. The silicon substrate layout is shown in figure 1(A), whereas figure 1(B) shows scanning electron microscopy images of three areas with characteristic pillar designs.

We first analyzed the influence of pillar dimension on SGN presence. We immunostained spiral ganglion neurons with Tuj1+ after seven days in culture, and described their position on the substrate by fluorescent microscopy. In order to do that, we calculated the cumulative percentages among 12 different samples from three independent experiments (N = 3), for a total of n = 906 neurons (an average of 72.5 ± 20 neurons/MPS) and plotted the results in figure 1(C). The surface plot of SGN on MPS (figure 1(C)) revealed that SGN did not display a uniform distribution to the surface, but rather prefers certain specific pillar geometries. Subsequently, we grouped the pillar features according to spacing (S) into bins. We chose to analyze pillar spacing among pillar width since previously it has been found that pillar spacing was more influential in changing neuronal behaviour than pillar width (Micholt et al 2013). The percentage of SGN in the bins was calculated for each experiment and averaged (N = 3). The results revealed that the cell presence was enhanced on pillars with spacing between 1.2 μm and 2.4 μm. The percentage of SGN grown on pillar spacings between 1.2 μm and 2.4 μm was doubled (42 ± 3%) compared to SGN found on all the other pillar dimensions (22 ± 3%, on S = 0.6–1.0 μm; 17 ± 3% on S = 3.5–5 μm and 19 ± 3% on S = 7–15 μm) as shown in figure 1(D).

3.2. Presence of SGN on MPS is comparable to control glass surfaces

Next, we assessed the overall growth characteristics of SGN on silicon micro-pillars compared to the conventional culturing substrates, i.e. borosilicate glass coverslips. We considered Tuj1+ SGN grown on MPS and we compared cell numbers with the control glass coverslips. The pictures in figure 2(A) show an overview of the SGN cell culture. From a visual examination, it can be observed that neurons and non-neural cells are present both in control and MPS surfaces, and that neurons grow neurites in both conditions. The graph in figure 2(B) shows that there were no quantitatively significant differences between the total numbers of Tuj1+ grown on MPS (1.2 ± 0.3 cells mm−2 equivalent to 72.5 ± 20 cells/sample, where sample area is 64 mm2; N = 3) and control glass surfaces (1.4 ± 0.4 cells mm−2, equivalent to 181.2 ± 72 cells/sample, where sample area is 132 mm2; N = 3). This indicates that structured silicon does not change the overall growth of SGN in culture.

Figure 2.

Figure 2. Effect of MPSs on SGN presence. (A). Fluorescent images of dissociated primary SGN cultures on control and chip substrates stained with the neuronal marker Tuj1+ (green) and the nuclear marker DAPI (blue). Scale bar is 100 μm. (B). SGN presence after 7DIV on control and chip substrates (N = 3). ns = not significant (one-way ANOVA).

Standard image High-resolution image

3.3. Specific micro-pillar spacing elicits directional SGN neurite orientation

In order to describe how pillar dimensions (width and spacing) influenced the growth directionality of neurites, SGN were stained with Tuj1+ and visualized (figure 3(A)). The orientation of the neurites on micro-pillar surfaces with different dimensions was analyzed using ImageJ software. FFT-Oval Profile analysis showed that the hexagonal pattern of the pillars had a strong influence on the directionality of neurite outgrowth for pillar spacings between 0.6 and 2.4 μm (figure 3(B)). On these specific micro-patterns, neurites preferentially oriented along three directional axes, spaced by 60° angle intervals (30°; 90°; 150°). The interaction of aligned neurites with pillars can be observed in SEM images (figure 3(C)). On the other hand, the topographic guidance was highly reduced for spacings of 3.5 to 5 μm, and completely lost for spacings above 7 μm, suggesting that there is a threshold value of inter-pillar spacing to elicit neurite contact guidance. We did not observe significant effects of the pillar width on neuronal outgrowth (data not shown), as was previously reported (Micholt et al 2013).

Figure 3.

Figure 3. Effect of pillar spacing on neurite alignment. (A). Representative fluorescent images of Tuj1+ SGN alignment on substrates with pillars of different dimensions. Scale bar is 100 μm. (B). Distribution of the direction of neurite growth around the preferred angles of the substrate topography. A total number of 36 SGN were analyzed (N = 3). The radial values range from 0–250 representing the intensity of pixels from FFT images along the same angle, while the angles are from 0 to 180 degrees. The values obtained by FFT-Oval Profile picture analysis were averaged across neurons in the same space bin and plotted. (C). Representative scanning electron microscopy (SEM) images of aligned SGN on a substrate with pillars 1 μm wide and 1.6 μm spaced. White arrows indicate neurites. Scale bar is 20 μm (left) and 5 μm (right).

Standard image High-resolution image

3.4. Specific micro-pillar spacing promotes longer SGN neurite growth

Next, we investigated whether pillar geometry had an effect on SGN neurite length. For this analysis the longest growth process of every SGN was considered, and their length was measured by submitting confocal images into ImageJ. After 1 DIV, pillars with both inter-spacings of 0.6–1.0 μm and 1.2–2.4 μm promoted SGN growth more significantly, as neurite length was significantly higher compared to the control group (figure 4). Among all pillar geometries, spacings of 1.2–2.4 μm gave rise to the longest neuritis; significantly longer than the control group. As expected, at 7 DIV in cultures, all neurites were longer than on the first day, but the increase was larger for the micro-structured surfaces, suggesting faster growth on these surfaces.

Figure 4.

Figure 4. Effect of pillar spacing on the longest neurite length. Box and whisker plots of the distribution of the longest neurite length as a function of pillar spacing and time. Bars represent the minimum and the maximum values of the data distribution, while the boxes are the interquartile range, and the central line indicates the median values. A total number of 302 SGN were analyzed (N = 3). Significant differences are indicated by # between 1DIV and 7DIV and by * to control.* = p < 0.05 and ** = p < 0.01 (one-way ANOVA).

Standard image High-resolution image

3.5. MPS induce less multipolar SGN morphology

We assessed the morphology of the SGN cells on both MPSs and control surfaces. The morphological shape of the SGN was determined by the number of neurites sprouting from the SGN soma, yielding neurite-free, monopolar, bipolar and multipolar morphologies (figure 5(A)). No significant differences were present between control and MPS surfaces concerning the percentage of neurite-free neurons (20 ± 5% and 30 ± 10% respectively). After 1 DIV, the MPS induced 3-fold more SGN with a monopolar morphology (46 ± 17%) with respect to the control condition (14 ± 8%) and 8-fold less SGN with a multipolar morphology (2 ± 4%) with respect to the control condition (17 ± 4%) (figure 5(B)). After 7 DIV, the neurite-free neurons on MPS decreased to 14 ± 6%. Monopolar neurons were still more abundant on MPS than in the control condition (39 ± 11% and 19 ± 10% respectively) while the number of multipolar SGN was significantly increased on both the MPS and control surfaces. However, the number of SGN with multipolar morphology was still lower on the MPS (14 ± 6%), almost half of that on the control surfaces (29 ± 12%) (figure 5(C)). In comparison with the directed growth of SGN on MPS as described above, these results revealed that MPS encouraged neurons to have less of a multipolar morphology.

Figure 5.

Figure 5. Effect of MPS on SGN neuronal morphology. (A). Examples of neurite-free, monopolar, bipolar and multipolar morphologies of Tuj1+ immunostained SGN. Scale bar is 50 μm. (B). Quantification of SGN neuronal morphology on control and chip substrates after 1DIV and (C). After 7DIV where N° indicates the number of sproutings (0 = neurite-free, 1 = monopolar, 2 = bipolar, 3+ = multipolar). A total number of 441 SGN were analyzed (N = 3). Significant differences are indicated by # to 1DIV and by * between groups.* = p < 0.05; ** = p < 0.01 (one-way ANOVA).

Standard image High-resolution image

3.6. MPSs decrease longest neurite-glial cells interactions

Previous observations of SGN in in vitro cell cultures have shown that these neurons often grow their axons in contact with S100+ cells (Bostrom et al 2007, Rak et al 2011). We observed that S100+ cells were the most abundant non-neural cell type in culture (60 ± 18%) and their number was stable over time both on control surfaces and MPS. S100+ cells showed heterogeneous morphologies, suggesting the presence of different types of glial cells, most probably Schwann cells and satellite cells (Whitlon et al 2009, Liu et al 2012). The ratio of glial cells/neurons was 12.5 ± 6.0, and there were no significant differences between the control surfaces and MPSs (Total number of SGN analyzed was n = 1205 from 16 images of MPS and seven images of control conditions). Further, we analyzed the interaction of the longest neurites from Tuj1+ cells with S100+ cells after 1 DIV and after 7 DIV on both surfaces (figure 6(A)).

Figure 6.

Figure 6. Effect of MPS on the longest neurite interaction with S100+ cells. (A). Examples of SGN cultures immunostained at 7DIV with the neuronal marker Tuj1+ (green), the glial marker S100 (red) and the nuclear marker DAPI (blue). Scale bar is 50 μm. (B). Quantification of the interaction of the longest neurite with S100+ glial cells on control and MPS as function of time. A total number of 235 SGN were analyzed (N = 3). Significant differences are indicated by # to 1DIV and by * between groups.* = p < 0.05; ** = p < 0.01 (one-way ANOVA).

Standard image High-resolution image

As can be seen in figure 6(B), the longest neurites on an MPS were encouraged to grow more in contact with micro-pillars than to seek interaction with glial cells that were present in their surroundings. After 1 DIV, on the control surfaces 46 ± 15% of the longest neurites sprouting from SGN were in contact with S100+ cells, while on the MPS the percentage of longest neurites touching glial cells was almost 4 times lower (9 ± 2%). After 7 DIV, in both conditions the number of longest neurites touching/contacting S100+ cells increased significantly. However, on MPS the number of longest neurites in contact with glial cells was still more than halved (33 ± 15%) compared to control surfaces (78 ± 25%). It seems that MPS present SGN with a strongly organized environment, which makes them reach out less to other supporting structures in their environment.

4. Discussion

Despite the success of cochlear implants, current technology is in need of further improvement to offer a more thorough solution to auditory-impaired people. The efficiency of the neuro-electronic interface remains the weak point of this technology. One likely approach would be to refine the electrode manufacturing process using micro-electrode arrays fabricated by silicon-based complementary metal–oxide–semiconductor (CMOS) technology, thereby increasing the number and density of stimulating electrodes, and integrating the stimulation circuitry into the chip. With this possible application in mind, in this study we tested the performance of silicon micro-pillar functionalized surfaces, as directly related to the design of high-resolution microelectrode arrays (MEAs). Organized and highly structured micro-patterned surfaces are not only an emergent potential tool to control cell growth in vitro, but could also provide strong evidence to foster an improved contact between neurons and electrodes in neuro-prostethic devices (Clarke et al 2011, Rak et al 2011).

Our results demonstrate the ability of MPS to support normal SGN neuron growth, enhance neurite alignment, promote outgrowth and induce dedicated morphologies on micro-pillars. First of all, MPS promoted a satisfactory SGN presence. Considering the number of seeded cells/sample (20 × 103), the calculated ratio of glial cells/neurons (12.5 ± 6; Results section 3.6) and the number of SGN/MPS (72.5 ± 20 neurons/sample), a survival rate of 4.8% is obtained for our experimental setup. When compared to other studies, this value can be considered satisfactory for primary SGN cell culture. For instance, Jin et al also report a rat P5 SGN survival rate of 0.98% after 3DIV (Jin et al 2013), Rak et al, who used neurons from P7 rats' cochlear nucleus, report a survival rate of 1.02% (Rak et al 2011), while Viera et al describe survival rate of adult SGN from rats, mice or guinea pigs ranging between 1 and 6% (Vieira et al 2007).

Spiral ganglion neurons turned out to exhibit enhanced growth when they adhered onto pillar arrays with a specific spacing range. The data obtained from neuronal presence, neurite growth and alignment together suggest that there is an optimal growth behaviour on pillars with spacing ranging from 0.6 to 2.4 μm. In particular, spacing between 1.2 and 2.4 μm promoted improved SGN growth and neurite elongation. Whether this preferential positioning is due to differential attachment or later survival selectivity is still an open question and is not addressed by this study. However, these trends of performance of specific geometries were confirmed by other studies. For instance, Micholt et al observed an optimal performance on embryonic hippocampal neurons of micro-pillars with widths and spacing between 1 and 2 μm, resulting in the formation of the longest neurite after 30 h of culture (Dowell-Mesfin et al 2004, Micholt et al 2013). Kundu et al proved that when pillars are in a specific range (1.4–1.8 μm spacing) the surface pattern was able to compete and overcome the repulsion of the soluble factor Semaphorin3A (Kundu et al 2013). Other examples are reviewed in Hoffman-Kim et al 2010. Taken together, these results, supported by other works, indicate that topographical cues strongly promote neuronal polarization and axonal guidance (Gomez et al 2007).

The presence of particular organizations of micro-pillars on silicon substrates also strongly influenced SGN morphology. Here we have shown that after 7 DIV, as an effect of time, neurons extended their processes 2.4 to 3 fold times longer than after 1 DIV, and developed more multipolar morphologies. However on MPSs, the monopolar and bipolar morphologies were more frequent, while multipolar morphologies developed significantly less than on the control surfaces. The early encouragement of different sprouting behaviors on MPS can be related to neuronal polarization and axonal path-finding. It was shown previously that hippocampal neurons at the edge of a flat and micro-pillar patterned surface preferentially orient toward the pattern, and sprout neurites in that direction (Micholt et al 2013). From this sprouting behaviour, it can be suspected that the morphology on structured surfaces will be altered toward less branched and thus more mono- and bipolar forms.

In vitro neurite-free, mono-, bi- and multi-polar morphologies have also been observed in other studies (Whitlon et al 2006, Whitlon et al 2007, Vieira et al 2007). The appearance of SGN in vitro may vary between studies, depending on the species of origin, the age of the animals and the cell culture protocol (Rusznak and Szucs 2009). In vivo SGN are bipolar cells that connect hair cells with the cochlear nucleus cells in the brainstem (Berglund and Ryugo 1987). In this study, the fact that MPSs discourage multipolar morphologies is indicative that these surfaces promote the morphology of SGN in vitro more similar to the in vivo original shape than for the control surface. The increase in bipolar neurons compared to monopolar neurons was promoted by BMP4 and LIF in a study of newborn mice SGN culture with a similar effect (Whitlon et al 2007). Moreover, a study by Khalifa et al (2013), reports that they observe a drastic increase of multipolar neurons from 5% on non-structured surfaces to 86–87% on surfaces with alternating lines. They state that 'the transition from a uni- or bipolar to a multipolar shape can happen as the neurons may not easily sense the chemical cues that are sorted in two opposite directions'. Something similar is probably happening on our MPS; however our rate of change into multipolar neurons doesn't seem to be drastically increased. Micro-topography probably acts on similar stimulation mechanisms that could be integrated synergistically with biochemical stimuli in the design of optimal electrode arrays.

SGN in vitro cultures also comprise an important presence of non-neuronal cells, mostly glial cells, where the ratio to neurons can range from 1:1 (Vieira et al 2007) to 20:1 (Rask-Andersen et al 2005). Schwann cells and satellite cells are the most commonly SGN-associated cells in the ganglion or spiral lamina, and they are immunoreactive for the marker S100 (Whitlon et al 2009, Liu et al 2012). In this study the ratio of S100+ cells to neurons was 12.5 ± 6.0:1. These types of glial cells play an important role in neuronal growth, survival, regeneration and axonal guidance in vivo and in vitro (Gillespie and Shepherd 2005, Whitlon et al 2009, Clarke et al 2011, Shibata et al (2011), Mattotti et al 2012, Provenzano et al 2011). It has been proven that SGN in culture prefer to grow close to glial cells, due to the adhesive molecules patterns, like Laminin-1, that these cells secrete (Whitlon et al 2009, Clarke et al 2011). Our data show that on MPSs an increased number of neurons extend their longest neurite in direct contact with the surface topography instead of seeking interactions with S100+ cells. Previous studies by Micholt et al, and others (Jang et al 2010, Ferrari et al 2010, Dowell-Mesfin et al 2004) demonstrated cytoskeletal alterations and upstream signalling related to neurons interacting with pillars and grooves, thus, our findings here are consistent with the bibliography. This suggests that topographic cues presented by the micro-pillars support direct neurite guidance without the mediating action of glial cells, creating a permissive environment for neuronal growth and neurite sprouting.

One of the approaches for the next generations of neuro-electronic interfaces for cochlear implants is the promotion of regrowth of peripheral SGN axons to enhance contact with the electrode array (Clarke et al 2011). Micro-structured surfaces such as the pillar arrays presented here could serve both for the integration of microelectrodes and for the enhancement of neurite guidance toward the implant to promote a cell-interface direct contact, mimicking the natural cochlear innervation of hair cells.

Previous work has demonstrated several in vitro applications of silicon micro-pillar electrodes integrated onto a CMOS chip for stimulation and recording of cells (Braeken et al 2009, Huys et al 2010, Huys et al 2012).

In conclusion, with this study we demonstrate the feasibility of MPS in supporting SGN growth: micro-pillars of a particular size-range (1.2–2.4 μm) were found to be optimal in promoting SGN presence, neurite growth and alignment. Moreover we demonstrated that MPS encourage more monopolar and multipolar SGN morphologies, and finally that MPS induce neurite growth with minimal interaction of S100+ glial cells.

Acknowledgments

This research was financially supported by grants from the Business Innovation Croatian Agency BICRO (PoC-IV-10) and the European Commission (FP7-CIG-2011-303927) awarded to DK. We would like to thank Professor Damir Sapunar for microscopy assistance, Jordi Cools for SEM images, Jagoda Dujić for the help with animal facility, Katarina Madirazza, Danijel Nejašmić, Josip Angelo Borovac and Adriana Banožić for the useful discussion about results and statistics.

Please wait… references are loading.
10.1088/1741-2560/12/2/026001