Supplementary Materials for - Science...2015/10/14  · Single-walled carbon nanotubes (CNTs) were...

27
www.sciencemag.org/content/350/6258/314/suppl/DC1 Supplementary Materials for A skin-inspired organic digital mechanoreceptor Benjamin C.-K. Tee, Alex Chortos, Andre Berndt, Amanda Kim Nguyen, Ariane Tom, Allister McGuire, Ziliang Carter Lin, Kevin Tien, Won-Gyu Bae, Huiliang Wang, Ping Mei, Ho-Hsiu Chou, Bianxiao Cui, Karl Deisseroth, Tse Nga Ng, Zhenan Bao* *Corresponding author. E-mail: [email protected] Published 16 October 2015, Science 350, 314 (2015) DOI: 10.1126/science.aaa9306 This PDF file includes: Materials and Methods Figs. S1 to S14 Tables S1 and S2 Caption for movie S1 References Other supplementary material for this manuscript includes the following: Movie S1

Transcript of Supplementary Materials for - Science...2015/10/14  · Single-walled carbon nanotubes (CNTs) were...

  • www.sciencemag.org/content/350/6258/314/suppl/DC1

    Supplementary Materials for

    A skin-inspired organic digital mechanoreceptor

    Benjamin C.-K. Tee, Alex Chortos, Andre Berndt, Amanda Kim Nguyen, Ariane Tom, Allister McGuire, Ziliang Carter Lin, Kevin Tien, Won-Gyu Bae, Huiliang Wang, Ping Mei,

    Ho-Hsiu Chou, Bianxiao Cui, Karl Deisseroth, Tse Nga Ng, Zhenan Bao* *Corresponding author. E-mail: [email protected]

    Published 16 October 2015, Science 350, 314 (2015) DOI: 10.1126/science.aaa9306

    This PDF file includes: Materials and Methods

    Figs. S1 to S14

    Tables S1 and S2

    Caption for movie S1

    References

    Other supplementary material for this manuscript includes the following: Movie S1

  • 2

    Materials and Methods

    Ring oscillator fabrication and testing

    The organic field-effect transistors are in top-gate geometry and are fabricated on plastic

    polyethylene naphthalate foils as described in Refs (15, 16). The electrodes are inkjet printed using

    Ag nanoparticle dispersion from Colloidal Ink in Japan. Organic semiconductors from Flexink and

    Polyera were inkjet printed to form p-channel and n-channel transistors. The gate dielectric is a

    bilayer, comprised of a Teflon layer (from Dupont) next to the semiconducting layer and a high-k

    PVDF-TrFE-CTFE relaxor polymer (from Piezotech). All fabrication and measurement are done

    in air. We tested the ring oscillator using an active probe and low-input capacitance unity gain

    buffer, which was used for interfacing a conventional silicon edge-detector chip with the organic

    circuits For optical neuron stimulation, an edge detector circuit was added to obtain a pulse width

    of 2 ms. With continuous stimulation, the neurons fire at their natural frequency (~300 Hz).

    Consequently, the pulse width must be smaller than the inverse of this frequency (3.3 ms) to avoid

    producing multiple action potentials with one oscillation of the oscillator. Therefore, the edge

    detector was necessary to ensure the desired property of producing one action potential per

    oscillation of the sensor. The complete circuit diagram is included as Fig. S14. The buffer and edge

    detector circuits are conventional silicon chips, but in future versions, it could be straightforward

    to eliminate the buffer and monolithically integrate the oscillator and edge detector on one substrate

    to have a fully printed organic digital e-skin.

    Piezoresistive sensor fabrication

    Single-walled carbon nanotubes (CNTs) were obtained from Carbon Solutions, Inc. (P2-SWNT).

    10 mg of CNTs were dispersed in NMP by sonicating for 30 minutes using a tip horn sonicator

    (Cole Parmer ultrasonic processor 750 W) at 30% power. The resulting solution was centrifuged

    at 8000 RPM for 30 minutes using a Sorvall Lynx 4000 centrifuge. The solution was spraycoated

    onto a silicon wafer at 200°C to form a conducting electrode. A solution of polyurethane

    elastomer (SG80A from Lubrizol) was spincoated onto the electrode. The electrode was

    subsequently transferred onto a polyurethane substrate. To form the piezoresistive composite, 20

    mg of CNTs in 9 mL of chloroform were dispersed by sonicating for 20 minutes at 30% power.

    After adding 1 mL of a 10 mg/mL solution of poly-3-hexylthiophene (Sepiolid P100), the

    solution was sonicated for another 60 minutes, and was subsequently centrifuged for 20 minutes

    at 6000 RPM. The dispersed CNTs were mixed with a solution of polyurethane (SG80A from

    Lubrizol) in chloroform and spincoated onto PDMS molds of pyramid structures. The CNT

    electrodes were laminated on top of the piezoresistive pyramids and heated at 60°C for 5 minutes

    to ensure good bonding. The CNT electrode/piezoresistive composite stack was then removed

    from the mold. As the counter electrodes, we used evaporated gold on glass slides as rigid

    electrodes or spraycoated silver nanowires (AW060 from Zhejiang Kechuang Advanced

    Materials Technology Co., Ltd.) embedded in PDMS as stretchable electrodes.

    Mutagenesis of iC1C2(s/v) and testing in cultured neurons

    DNA sequence of n-terminal modified C1C2 version bC1C2 was synthesized (Genescript) and

    cloned into AAV vectors containing the CamKIIα promoter for expression in neurons. The

    construct was fused to eYFP DNA (enhanced yellow fluorescent protein) to detect protein

    expression in neurons by fluorescence microscopy. Mutations (I170V and Q95S) were introduced

    using QuickChangeTM Site-Directed mutagenesis kit (Agilent). Plasmid DNA was purified with

    QIAprepTM Spin Miniprep Kits (Qiagen) after transformation and amplification in E. coli.

  • 3

    Electrophysiological recordings in neuronal cultures were prepared as described(27). Patch pipettes

    (4-6 MΩ) were pulled from glass capillaries (Sutter Instruments) with a horizontal puller (P-2000,

    Sutter Instruments) for whole-cell recordings in voltage and current clamp. Recordings were made

    using a MultiClamp700B amplifier (Molecular Devices). The external recording solution contained

    (in mM): 127 NaCl, 10 KCl, 10 HEPES, 2 CaCl2, 2 MgCl2, 30 D-glucose, pH 7.3, including

    synaptic blockers (25 µM D-APV, 10 µM NBQX). The patch pipette solution contained (in mM):

    140 K-gluconate, 10 HEPES, 10 EGTA, 2 MgCl2, pH 7.3. Series resistance was monitored

    throughout recordings for stability. A Spectra X Light engine (Lumencor) was used to excite eYFP

    and to apply light for opsin activation. Stimulation light (475 nm) was applied through a 40X

    objective (Olympus) at 5 mW/mm2 light intensity. Photocurrents were measured at -80 mV in

    voltage clamp mode during 1 s light application. All reported values refer to the stationary

    photocurrents. The activation spectra for bC1C2(s/v) and ChR2(HR) were measured by recording

    stationary photocurrent in voltage clamp mode at -80 mV and light intensities of 0.65 mW/mm2 at

    each wavelength. Light was delivered through 20 nm bandbass filters (Thorlabs) at (in nm): 400,

    420, 440, 460, 480, 500, 520, 540, 560, 580, 600. Photocurrents were normalized to 480 nm.

    Kinetics of channel closure were determined by fitting the decay of photocurrents after light-off,

    with mono-exponential or bi-exponential functions. Channel kinetics were quantified by

    corresponding tauoff values respectively. pClamp10.3 (Molecular Devices) and OriginLab8

    (OriginLab) software was used to record and analyze data.

    Stereotactic virus injection

    The following adeno-associated viruses (AAV) with serotype DJ were produced at the Stanford

    Neuroscience Gene Vector and Virus Core: double floxed AAVdj-EF1a-DIO::ChR2(E123A)-

    EYFP and AAVdj-EF1a-DIO::bC1C2(s/v)-EYFP 4-6 week old PV+-IRES-Cre mice were injected

    bilaterally with 1 µl of either virus in the medial prefrontal cortex (mPFC) or the primary

    somatsensory cortex forelimb region (S1FL), at the following coordinates (from bregma): AP +1.7

    mm, ML +0.3 mm, DV -2.5 mm (mPFC) or AP +0.26 mm, ML +2.1 mm, DV -1.5 mm (S1FL).

    Titer were matched at 1.5x1012 vg/ml for both viruses. Expression is controlled by Cre-recombinase

    and restricted to parvalbumin positive cells (PV+). Expression is controlled by Cre-recombinase

    and restricted to parvalbumin positive cells (PV+).

    Acute slice electrophysiology

    Acute brain slices were prepared from mouse medial prefrontal cortex (mPFC) or primary

    somatsensory cortex forelimb region (S1FL) 4-6 weeks after viral injection. Coronal slices (300

    μm) from injected mice were prepared after intracardial perfusion with ice-cold, sucrose-containing

    artificial cerebrospinal fluid solution (ACSF; in mM): 85 NaCl, 75 sucrose, 2.5 KCl, 25 glucose,

    1.25 NaH2PO4, 4 MgCl2, 0.5 CaCl2 and 24 NaHCO3. Slices recovered for 1 hour at 32–34 °C, and

    then were transferred to an oxygenated recording ACSF solution (in mM): 123 NaCl, 3 KCl, 26

    NaHCO3, 2 CaCl2, 1 MgCl2, 1.25 NaH2PO4 and 11 glucose, at room temperature.

    Electrophysiological recordings were performed at 32–34 °C under constant perfusion of the

    oxygenated recording ACSF solution. For mPFC recordings, synaptic transmission blockers (D-2-

    amino-5-phosphonovaleric acid (APV; 25 μM), 2,3-dihydroxy-6-nitro-7-sulfamoyl-

    benzo[f]quinoxaline-2,3-dione (NBQX; 10 μM) and gabazine (10 μM)) were added to the

    recording ACSF solution. Slices were visualized with an upright microscope (BX61WI, Olympus)

    under infrared differential interference contrast (IR-DIC) optics. A Spectra X Light engine

    (Lumencor) was used both for viewing fluorescent protein expression and delivering light pulses

    for opsin activation. PV+ neurons were identified by eYFP+ expression. Whole-cell voltage-clamp

    recordings were performed at -80 mV, and current-clamp recordings were performed at rest. Patch-

    clamp pipettes contained the following internal solution (in mM): 125 K-gluconate, 10 KCl, 10

    HEPES, 4 Mg3-ATP, 0.3 Na-GTP, 10 phosphocreatine, 1 EGTA. Recordings were conducted using

    MultiClamp700B amplifier and pClamp10.3 software (Molecular Devices). pClamp10.3,

  • 4

    OriginLab8 (OriginLab), and SigmaPlot (SPSS) were used to analyze data. Stationary photocurrent

    of the opsins was measured at the end of a 1-s light pulse (475 nm) in voltage-clamp mode. Light-

    evoked spike probability in the PV+ neurons was calculated as the fraction of successful action

    potentials evoked in neurons at light stimulation frequencies ranging from 40 to 200 Hz during a 4

    second stimulation period.

    Electrical neuron stimulation with inorganic oscillators

    Prior to cell plating, multi-electrode arrays (MEAs) were cleaned with a brief oxygen plasma and

    coated with 200 μg/mL poly-L-lysine. Hippocampal [or cortical] neurons were isolated from E18

    rats and cultured for 10-14 days on MEAs as previously described. The neurons were maintained

    in neurobasal medium supplemented with B27, penicillin/streptomycin, and L-glutamine.

    Electrical activity of primary neuron cultures was stimulated by an external stimulation

    macroelectrode (to achieve spatial control) and recorded by planar (Multi-electrode arrays) MEAs.

    The stimulating macroelectrode was connected to a pulse generator, which was in turn connected

    to the DiTact sensor. Commercial signal processing and amplification were used (Multichannel

    Systems) to record both stimuli and evoked cellular responses.

  • 5

    DiTact sensor characterization

    Fig. S1. Images of organic ring oscillators. (A) Array of printed complementary ring oscillators.

    Zoomed-in area shows the 3-stage ring oscillator. Each oscillator pixel is composed of three

    inverters that form the oscillator and one additional inverter that functions as a buffer. The

    oscillators near the top of the substrate are five stage oscillators and the ones near the bottom of the

    substrate are three-stage oscillators (B) Voltage output of a three-stage ring oscillator showing a

    frequency of 110 Hz at a supply voltage of 8V.

  • 6

    Fig. S2. Electrical characteristics of p- and n-type transistors in the complementary organic

    oscillators. (A) Transfer curves and (B) output curves for p-type devices. P-type devices had a W/L

    of 42 and a mobility of 0.10 cm2V-1s-1. (C) Transfer curves and (D) output curves for n-type devices.

    N-type devices had a W/L of 60 and a mobility of 0.07 cm2V-1s-1.

  • 7

    Pressure Sensor Characterization

    Fig. S3. Effect of pyramid size on the pressure response of piezoresistive sensors. Small

    pyramids (3 μm) have higher sensitivity, but smaller dynamic range. 50 μm pyramids have the

    desired sensing properties throughout the pressure range of interest. The voltage was 5 V.

  • 8

    Fig. S4. Effect of pyramid spacing on the pressure response of the piezoresistive sensors. As

    the spacing between pyramids increases, the sensitivity increases. However, when pyramids are

    spaced far apart, the region between the pyramids comes in contact with the substrate, preventing

    recovery to the initial impedance. The pressure at which device characteristics become

    irrecoverable decreases as pyramid spacing increases. With a spacing of 50 μm, the device

    properties are recoverable throughout the measurement range. With a spacing of 100 μm,

    irrecoverable deformation occurs at ~40 kPa, and with a spacing of 200 μm, irrecoverable

    deformation occurs at only 20 kPa. The voltage was 5 V.

    10

    Figure S4. Effect of pyramid spacing on the pressure response of the piezoresistive sensors. As the

    spacing between pyramids increases, the sensitivity increases. However, when pyramids are spaced

    far apart, the region between the pyramids comes in contact with the substrate, preventing recovery

    to the initial impedance. The pressure at which device characteristics become irrecoverable

    decreases as pyramid spacing increases. With a spacing of 50 μm, the device properties are

    recoverable throughout the measurement range. With a spacing of 100 μm, irrecoverable

    deformation occurs at ~40 kPa, and with a spacing of 200 μm, irrecoverable deformation occurs at

    only 20 kPa. The voltage was 5 V.

  • 9

    Fig. S5. Voltage dependence of the sensor impedance. The voltage dependence can be

    approximated as exponential (R2 > 0.98).

    Data Fitting for Sensor Pressure Dependence

    The sensors function partially based on changes in the tunneling barrier as a function of

    pressure. Tunneling current density is substantially influenced by the applied voltage.

    Consequently, the pressure sensor characteristics are voltage-dependent. This is important for the

    DiTact sensor because in a voltage divider configuration, the voltage across the piezoresistor

    changes with applied pressure, resulting in a complex relationship between the pressure and voltage

    dependence. Fig. S6 depicts the voltage dependence and fitted characteristics of a model sensor.

  • 10

    Fig. S6: Experimental and fitted sensor impedance. (A) Dependence of the impedance of the

    sensor on the pressure and voltage. (B) Fitted value of the impedance as a function of pressure and

    voltage. The fitting process gave an R2 value of 0.975. (C) The log value of the residuals from the

    fitting process.

    The magnitude of the impedance can be modeled using equation (1), where Zsensor is the

    impedance of the sensor, P is the pressure, Vsensor is the voltage across the sensor, and a1 to a5 are

    parameters to be fit to the data. The electric field dependence of tunneling currents from CNT

    electrodes can often be approximated by an exponential fit(28, 29). The impedance of a

    piezoresistive sensor based on tunneling depends exponentially on the applied strain(30).

    Consequently, the pressure dependence is determined by the stress-strain behavior of the device.

  • 11

    The deviation from exponential behavior caused by the shape of the pyramids is taken into account

    by the exponent on the pressure term (a4).

    𝑍𝑠𝑒𝑛𝑠𝑜𝑟 = 𝑎110^(𝑎2𝑉𝑠𝑒𝑛𝑠𝑜𝑟 + 𝑎3𝑃𝑎4) (1)

    The circuit diagram for the sensing element is show in Fig. S7. The reference resistor is not

    necessary in the device, but can be useful for tuning the pressure response characteristics (Fig. S9).

    Therefore, in addition to the piezoresistive composite composition and the structure of the pyramids

    in the piezoresistive sensor, the reference resistor adds another method of modifying the device

    output. Tunable characteristics are important because mechanoreceptors in different parts of the

    body have different saturation frequencies, different sensitivities, and different shapes of the

    frequency vs stimulus curves(18).

    Equation (2) provides the voltage across the oscillator which is the supply voltage of the

    oscillator. The fitted equation for the impedance of the sensor (1) can be substituted into the

    equation for the voltage across the oscillator (2), resulting in a non-linear implicit equation for the

    dependence of Vosc on the applied pressure. Furthermore, as shown in Fig. 1c, the impedance

    through the oscillator is also voltage-dependent. The resulting equation was solved using MATLAB

    to determine the voltage across the sensor and the oscillator for each pressure value. The pressure

    and the voltage across the sensor and the impedance of the sensor for the fitted data from Fig. 2c

    are plotted in Fig. S8.

    𝑉𝑜𝑠𝑐 =[𝑍𝑜𝑠𝑐

    −1 +𝑍𝑟𝑒𝑠𝑖𝑠𝑡𝑜𝑟−1 ]

    −1

    [𝑍𝑜𝑠𝑐−1 +𝑍𝑟𝑒𝑠𝑖𝑠𝑡𝑜𝑟

    −1 ]−1

    + 𝑍𝑠𝑒𝑛𝑠𝑜𝑟𝑉𝑎𝑝𝑝𝑙𝑖𝑒𝑑 (2)

  • 12

    Fig. S7. Circuit diagram used for calculating pressure dependence. The voltage dependence of

    the pressure sensor characteristics as well as the voltage dependence of the oscillator impedance

    must be taken into account to accurately predict the frequency of the sensor.

  • 13

    Fig. S8. Sensor voltage fit during pressure test. Fitted values of the impedance of the pressure

    sensor and the voltage across the sensor as a function of pressure for the circuit shown in fig. S7

    without a resistor (resistor with infinite impedance).

    The modeled electrical properties of the circuit was used to investigate some aspects of the circuit’s

    response to pressure. Fig. S6 shows the effect of different reference resistor values on the pressure

    dependence of the sensing characteristics.

  • 14

    Fig. S9: Effect of reference resistors on frequency output. As the impedance of the reference

    resistor decreases, the sensitivity decreases while the working range increases. Furthermore, the

    shape of the curve becomes more linear.

  • 15

    Channelrhodopsin Protein Characterization

    We replaced the first 49 amino acids by the first 11 amino acids of ChR2 to improve membrane

    trafficking and expression. As shown before, introduction of the mutation I131V accelerated

    channel closure(31), which we further enhanced by replacing glutamine 56 by serine (Q56S).

    Fig. S10. bC1C2(v/s) characterization in cultured neurons. (A) Schematic composition of

    channelrhodopsin hybrids C1C2 and bC1C2(v/s): Both contain segments from channelrhodopsin 1

    and 2 (ChR1 and 2). Numbers indicate amino acid positions. bC1C2(s/v) was mutated at positions

    56 (Q56S) and 131 (I131V) to accelerate channel kinetics. The first 39 amino acids were replaced

    by the first 11 amino acids of ChR2 to improve expression and membrane trafficking. (B)

    Activation spectrum of bC1C2(v/s) (red) compared to ChR2(HR) (black). Each construct n = 3. All

    error bars indicate s.e.m., n indicates number of cells. (C) Photocurrents amplitudes -80 mV: ChR2-

    H134R (ChR2(HR), n = 24), ChR2-E123A (ChETA(EA), n = 18), C1C2 (n = 18), bC1C2(IV)

    (bC1C2(I131V), n = 15) and bC1C2(s/v) (n = 19). (D) Left: tau-values of mono-exponential off-

    kinetics (ChR2(HR), ChETA(EA), C1C2) and the fast components of bi-exponential off-kinetics

    (C1C2(IV), bC1C2(s/v)). Same n as in (A). Right: Proportion of photocurrent amplitudes decaying

    with fast tau-off.

  • 16

  • 17

    Fig. S11. Comparison of fast channelrhodopsin bC1C2(v/s) and ChETA(EA) (A) Decay of

    photocurrents after 475 nm light excitation (blue bar). Current traces were normalized to compare

    speed of channel closures. bC1C2(s/v) (red) off-kinetics are bi-exponential with a fast and slow

    component. ChETA(EA) (black) current decays exponentially. (B) Photocurrent amplitudes

    measured in cortical parvalbumin positive (PV+), fast spiking interneurons of mice medial

    prefrontal cortex (mPFC). ChETA(EA): n = 16, bC1C2(s/v): n = 10. (C) Left: tau-values of mono-

    exponential ChETA(EA) off-kinetics and the fast bi-exponential bC1C2(s/v) off-kinetics.

    ChETA(EA): n = 16, bC1C2(s/v): n = 10. Right: Proportion of photocurrent amplitudes decaying

    with fast and slow tau-off.

  • 18

    Fig. S12. Comparison of real and artificial mechanoreceptors. (A) Pressure response of

    biological mechanoreceptors in the human foot.(32) Reproduced with permission. (B) Frequency

    of action potentials vs pressure for parvalbumin interneurons in the prefrontal cortex coupled to the

    DiTact sensor. There is a good correlation in the threshold frequency and maximum frequency. The

    first three trials depicted were from different locations on a sample with a CNT content of 1.17

    wt%. The fourth trial was conducted using a sensor with a smaller concentration of CNTs and

    shows a lower sensitivity. The operation voltage was 8 V, resulting in a max frequency of 130 Hz

    and a max power consumption of 4.5 µW. (C) Indentation response of biological mechanoreceptors

    in the human fingertip.(33) Reproduced with permission. (D) Frequency of action potentials vs

    pressure for neurons in the primary somatosensory cortex coupled to the DiTact sensor. All ten

    trials used the same oscillator and sensor. The operation voltage was 11 V, resulting in a max

  • 19

    frequency of 200 Hz and a max power consumption of 18.4 µW.

    Fig. S13. Electrical stimulation of plated hippocampal neurons.. (A) A ramped pressure from 0

    to 100kPa was applied. The stimulation pulse frequency increases proportionally with the pressure

    with the neuron firing tracking the stimulation frequency. (B) Three instance of stimulated neuron

    firing from tactile sensor. (C) Panel shows a single action potential triggered from a stimulation

    pulse due to pressure. An inorganic oscillator was used as the source of oscillation.

  • 20

    Fig. S14: Full circuit diagram of the elements used for interfacing the DiTact sensor with the neuron

    stimulation equipment. The buffer is a unity gain low-input capacitance node for interfacing

    conventional silicon edge-detector chip with organic circuits.

  • 21

    Photocurrents

    (pA)

    ± s.e.m / n

    tauoff

    (ms)

    ± s.e.m /

    n

    fast tauoff

    (ms)

    ± s.e.m / n

    slow tauoff

    (ms)

    ± s.e.m

    % of

    photocurrent

    w/ fast tauoff

    ± s.e.m

    % of

    photocurrent

    w/ slow tauoff

    ± s.e.m

    ChR(HR) 868.8

    ± 79.4 / 24

    38.8

    ± 8.7 / 24

    - - - -

    ChETA(EA) 657.9

    ± 72.0 / 18

    18.8

    ± 1.3 / 18

    - - - -

    C1C2 1212.9

    ± 129.9 / 18

    30.7

    ± 1.0 / 18

    - - - -

    bC1C2(IV) 1284.5

    ± 98.7 / 15

    - 14.7

    ± 0.81 / 15

    137.8

    ± 14.9

    88.2

    ± 0.1

    11.8

    ± 0.1

    bC1C2(s/v) 1245.8

    ± 118.87 / 19

    - 8.9

    ± 0.55 / 19

    129.7

    ± 20.8

    87.0

    ± 0.1

    13.0

    ± 0.1

    Supp. Table 1. Biophysical properties of channelrhodopsin variants in cultured neurons of

    mouse hippocampus. s.e.m. is standard error of the mean. n indicates number of cells

  • 22

    Photocurrents

    (pA)

    ± s.e.m / n

    tauoff

    (ms)

    ± s.e.m / n

    fast tauoff

    (ms)

    ± s.e.m / n

    slow

    tauoff

    (ms)

    ± s.e.m

    % of

    photocurrent

    w/ fast tauoff

    ± s.e.m

    % of

    photocurrent

    w/ slow tauoff

    ± s.e.m

    ChETA(EA) 311.1

    ± 50.2 / 16

    7.0

    ± 0.3 / 16

    - - - -

    bC1C2(s/v) 704.1

    ± 101.1 / 10

    - 3.4

    ± 0.2 / 10

    31.2

    ± 2.0

    78.6

    ± 1.6

    21.4

    ± 1.6

    Supp. Table 2. Biophysical properties of ChETA(EA) and bC1C2(s/v) in cultured neurons

    of mouse hippocampus. s.e.m. is standard error of the mean. n indicates number of cells.

  • 23

    Movie S1

    The movie shows the flexible DiTact sensors mounted on a wearable glove. At zero pressure stimuli, the sensor outputs no digital signal. The glove was used to apply pressure on a digital weighing scale. The output from the monitor shows the frequency output from the DiTact sensors. As the pressure increases, the frequency increase.

  • References

    1. L. R. Hochberg, D. Bacher, B. Jarosiewicz, N. Y. Masse, J. D. Simeral, J. Vogel, S. Haddadin,

    J. Liu, S. S. Cash, P. van der Smagt, J. P. Donoghue, Reach and grasp by people with

    tetraplegia using a neurally controlled robotic arm. Nature 485, 372–375 (2012). Medline

    doi:10.1038/nature11076

    2. J. E. O’Doherty, M. A. Lebedev, P. J. Ifft, K. Z. Zhuang, S. Shokur, H. Bleuler, M. A.

    Nicolelis, Active tactile exploration using a brain-machine-brain interface. Nature 479,

    228–231 (2011). Medline doi:10.1038/nature10489

    3. A. Abbott, Neuroprosthetics: In search of the sixth sense. Nature 442, 125–127 (2006).

    Medline doi:10.1038/442125a

    4. S. Raspopovic, M. Capogrosso, F. M. Petrini, M. Bonizzato, J. Rigosa, G. Di Pino, J.

    Carpaneto, M. Controzzi, T. Boretius, E. Fernandez, G. Granata, C. M. Oddo, L. Citi, A.

    L. Ciancio, C. Cipriani, M. C. Carrozza, W. Jensen, E. Guglielmelli, T. Stieglitz, P. M.

    Rossini, S. Micera, Restoring natural sensory feedback in real-time bidirectional hand

    prostheses. Sci. Transl. Med. 6, 222ra19 (2014). Medline

    doi:10.1126/scitranslmed.3006820

    5. M. Lotze, W. Grodd, N. Birbaumer, M. Erb, E. Huse, H. Flor, Does use of a myoelectric

    prosthesis prevent cortical reorganization and phantom limb pain? Nat. Neurosci. 2, 501–

    502 (1999). Medline doi:10.1038/9145

    6. D. W. Tan, M. A. Schiefer, M. W. Keith, J. R. Anderson, J. Tyler, D. J. Tyler, A neural

    interface provides long-term stable natural touch perception. Sci. Transl. Med. 6,

    257ra138 (2014). Medline doi:10.1126/scitranslmed.3008669

    7. S. C. B. Mannsfeld, B. C. Tee, R. M. Stoltenberg, C. V. Chen, S. Barman, B. V. Muir, A. N.

    Sokolov, C. Reese, Z. Bao, Highly sensitive flexible pressure sensors with

    microstructured rubber dielectric layers. Nat. Mater. 9, 859–864 (2010). Medline

    doi:10.1038/nmat2834

    8. G. Schwartz, B. C. Tee, J. Mei, A. L. Appleton, H. Kim, H. Wang, Z. Bao, Flexible polymer

    transistors with high pressure sensitivity for application in electronic skin and health

    monitoring. Nat. Commun. 4, 1859 (2013). Medline doi:10.1038/ncomms2832

    http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=22596161&dopt=Abstracthttp://dx.doi.org/10.1038/nature11076http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=21976021&dopt=Abstracthttp://dx.doi.org/10.1038/nature10489http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16837993&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=16837993&dopt=Abstracthttp://dx.doi.org/10.1038/442125ahttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=24500407&dopt=Abstracthttp://dx.doi.org/10.1126/scitranslmed.3006820http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=10448212&dopt=Abstracthttp://dx.doi.org/10.1038/9145http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25298320&dopt=Abstracthttp://dx.doi.org/10.1126/scitranslmed.3008669http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20835231&dopt=Abstracthttp://dx.doi.org/10.1038/nmat2834http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=23673644&dopt=Abstracthttp://dx.doi.org/10.1038/ncomms2832

  • 9. M. Kaltenbrunner, T. Sekitani, J. Reeder, T. Yokota, K. Kuribara, T. Tokuhara, M. Drack, R.

    Schwödiauer, I. Graz, S. Bauer-Gogonea, S. Bauer, T. Someya, An ultra-lightweight

    design for imperceptible plastic electronics. Nature 499, 458–463 (2013). Medline

    doi:10.1038/nature12314

    10. T. Sekitani, T. Yokota, U. Zschieschang, H. Klauk, S. Bauer, K. Takeuchi, M. Takamiya, T.

    Sakurai, T. Someya, Organic nonvolatile memory transistors for flexible sensor arrays.

    Science 326, 1516–1519 (2009). Medline

    11. D.-H. Kim, N. Lu, R. Ma, Y. S. Kim, R. H. Kim, S. Wang, J. Wu, S. M. Won, H. Tao, A.

    Islam, K. J. Yu, T. I. Kim, R. Chowdhury, M. Ying, L. Xu, M. Li, H. J. Chung, H. Keum,

    M. McCormick, P. Liu, Y. W. Zhang, F. G. Omenetto, Y. Huang, T. Coleman, J. A.

    Rogers, Epidermal electronics. Science 333, 838–843 (2011). Medline

    doi:10.1126/science.1206157

    12. J. Kim, M. Lee, H. J. Shim, R. Ghaffari, H. R. Cho, D. Son, Y. H. Jung, M. Soh, C. Choi, S.

    Jung, K. Chu, D. Jeon, S. T. Lee, J. H. Kim, S. H. Choi, T. Hyeon, D. H. Kim,

    Stretchable silicon nanoribbon electronics for skin prosthesis. Nat. Commun. 5, 5747

    (2014). Medline doi:10.1038/ncomms6747

    13. K. Takei, T. Takahashi, J. C. Ho, H. Ko, A. G. Gillies, P. W. Leu, R. S. Fearing, A. Javey,

    Nanowire active-matrix circuitry for low-voltage macroscale artificial skin. Nat. Mater.

    9, 821–826 (2010). Medline doi:10.1038/nmat2835

    14. E. P. Gardner, Touch (eLS/Wiley Online Library, 2001). doi:10.1038/npg.els.0000219

    15. T. N. Ng, D. E. Schwartz, L. L. Lavery, G. L. Whiting, B. Russo, B. Krusor, J. Veres, P.

    Bröms, L. Herlogsson, N. Alam, O. Hagel, J. Nilsson, C. Karlsson, Scalable printed

    electronics: An organic decoder addressing ferroelectric non-volatile memory. Sci. Rep.

    2, 585 (2012). doi:10.1038/srep00585

    16. P. Mei, T. N. Ng, R. A. Lujan, D. E. Schwartz, S. Kor, B. S. Krusor, J. Veres, Utilizing high

    resolution and reconfigurable patterns in combination with inkjet printing to produce high

    performance circuits. Appl. Phys. Lett. 105, 123301 (2014). doi:10.1063/1.4896547

    http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=23887430&dopt=Abstracthttp://dx.doi.org/10.1038/nature12314http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20007895&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=21836009&dopt=Abstracthttp://dx.doi.org/10.1126/science.1206157http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25490072&dopt=Abstracthttp://dx.doi.org/10.1038/ncomms6747http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20835235&dopt=Abstracthttp://dx.doi.org/10.1038/nmat2835http://dx.doi.org/10.1063/1.4896547

  • 17. T. Sekitani, U. Zschieschang, H. Klauk, T. Someya, Flexible organic transistors and circuits

    with extreme bending stability. Nat. Mater. 9, 1015–1022 (2010). Medline

    doi:10.1038/nmat2896

    18. M. Knibestöl, Stimulus-response functions of slowly adapting mechanoreceptors in the

    human glabrous skin area. J. Physiol. 245, 63–80 (1975). Medline

    doi:10.1113/jphysiol.1975.sp010835

    19. D. A. Nowak, J. Hermsdörfer, Grip force behavior during object manipulation in

    neurological disorders: Toward an objective evaluation of manual performance deficits.

    Mov. Disord. 20, 11–25 (2005). Medline doi:10.1002/mds.20299

    20. B. C. K. Tee, A. Chortos, R. R. Dunn, G. Schwartz, E. Eason, Z. Bao, Tunable flexible

    pressure sensors using microstructured elastomer geometries for intuitive electronics.

    Adv. Funct. Mater. 24, 5427–5434 (2014). doi:10.1002/adfm.201400712

    21. A. Canales, X. Jia, U. P. Froriep, R. A. Koppes, C. M. Tringides, J. Selvidge, C. Lu, C. Hou,

    L. Wei, Y. Fink, P. Anikeeva, Multifunctional fibers for simultaneous optical, electrical

    and chemical interrogation of neural circuits in vivo. Nat. Biotechnol. 33, 277–284

    (2015). Medline doi:10.1038/nbt.3093

    22. I. R. Minev, P. Musienko, A. Hirsch, Q. Barraud, N. Wenger, E. M. Moraud, J. Gandar, M.

    Capogrosso, T. Milekovic, L. Asboth, R. F. Torres, N. Vachicouras, Q. Liu, N. Pavlova,

    S. Duis, A. Larmagnac, J. Vörös, S. Micera, Z. Suo, G. Courtine, S. P. Lacour, Electronic

    dura mater for long-term multimodal neural interfaces. Science 347, 159–163 (2015).

    Medline doi:10.1126/science.1260318

    23. L. A. Gunaydin, O. Yizhar, A. Berndt, V. S. Sohal, K. Deisseroth, P. Hegemann, Ultrafast

    optogenetic control. Nat. Neurosci. 13, 387–392 (2010). Medline doi:10.1038/nn.2495

    24. J. Mattis, K. M. Tye, E. A. Ferenczi, C. Ramakrishnan, D. J. O’Shea, R. Prakash, L. A.

    Gunaydin, M. Hyun, L. E. Fenno, V. Gradinaru, O. Yizhar, K. Deisseroth, Principles for

    applying optogenetic tools derived from direct comparative analysis of microbial opsins.

    Nat. Methods 9, 159–172 (2012). Medline doi:10.1038/nmeth.1808

    25. H. E. Kato, F. Zhang, O. Yizhar, C. Ramakrishnan, T. Nishizawa, K. Hirata, J. Ito, Y. Aita,

    T. Tsukazaki, S. Hayashi, P. Hegemann, A. D. Maturana, R. Ishitani, K. Deisseroth, O.

    http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=21057499&dopt=Abstracthttp://dx.doi.org/10.1038/nmat2896http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=1127614&dopt=Abstracthttp://dx.doi.org/10.1113/jphysiol.1975.sp010835http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=15455447&dopt=Abstracthttp://dx.doi.org/10.1002/mds.20299http://dx.doi.org/10.1002/adfm.201400712http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25599177&dopt=Abstracthttp://dx.doi.org/10.1038/nbt.3093http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25574019&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=25574019&dopt=Abstracthttp://dx.doi.org/10.1126/science.1260318http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20081849&dopt=Abstracthttp://dx.doi.org/10.1038/nn.2495http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=22179551&dopt=Abstracthttp://dx.doi.org/10.1038/nmeth.1808

  • Nureki, Crystal structure of the channelrhodopsin light-gated cation channel. Nature 482,

    369–374 (2012). Medline

    26. H. Haeberle, E. A. Lumpkin, Merkel cells in somatosensation. Chemosens. Percept. 1, 110–

    118 (2008). Medline doi:10.1007/s12078-008-9012-6

    27. A. Berndt, S. Y. Lee, C. Ramakrishnan, K. Deisseroth, Structure-guided transformation of

    channelrhodopsin into a light-activated chloride channel. Science 344, 420–424 (2014).

    Medline doi:10.1126/science.1252367

    28. J.-M. Bonard, H. Kind, T. Stöckli, L.-O. Nilsson, Field emission from carbon nanotubes: The

    first five years. Solid State Electron. 45, 893–914 (2001). doi:10.1016/S0038-

    1101(00)00213-6

    29. M. A. McCarthy, B. Liu, A. G. Rinzler, High current, low voltage carbon nanotube enabled

    vertical organic field effect transistors. Nano Lett. 10, 3467–3472 (2010). Medline

    doi:10.1021/nl101589x

    30. D. Bloor, K. Donnelly, P. J. Hands, P. Laughlin, D. Lussey, A metal-polymer composite with

    unusual properties. J. Phys. D 38, 2851–2860 (2005). doi:10.1088/0022-3727/38/16/018

    31. J. Y. Lin, M. Z. Lin, P. Steinbach, R. Y. Tsien, Characterization of engineered

    channelrhodopsin variants with improved properties and kinetics. Biophys. J. 96, 1803–

    1814 (2009). Medline doi:10.1016/j.bpj.2008.11.034

    32. J. P. Vedel, J. P. Roll, Response to pressure and vibration of slowly adapting cutaneous

    mechanoreceptors in the human foot. Neurosci. Lett. 34, 289–294 (1982). Medline

    doi:10.1016/0304-3940(82)90190-2

    33. P. R. Burgess, J. Mei, R. P. Tuckett, K. W. Horch, C. M. Ballinger, D. A. Poulos, The neural

    signal for skin indentation depth. I. Changing indentations. J. Neurosci. 3, 1572–1585

    (1983). Medline

    http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=22266941&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=19834574&dopt=Abstracthttp://dx.doi.org/10.1007/s12078-008-9012-6http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=24763591&dopt=Abstracthttp://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=24763591&dopt=Abstracthttp://dx.doi.org/10.1126/science.1252367http://dx.doi.org/10.1016/S0038-1101(00)00213-6http://dx.doi.org/10.1016/S0038-1101(00)00213-6http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=20707327&dopt=Abstracthttp://dx.doi.org/10.1021/nl101589xhttp://dx.doi.org/10.1088/0022-3727/38/16/018http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=19254539&dopt=Abstracthttp://dx.doi.org/10.1016/j.bpj.2008.11.034http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=6298676&dopt=Abstracthttp://dx.doi.org/10.1016/0304-3940(82)90190-2http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&list_uids=6875657&dopt=Abstract

    SOM.page.1.EXPRESS.no.movies.pdfA skin-inspired organic digital mechanoreceptor

    aaa9306TeeRefs.pdfReferences