Science Education
>

Experimental Methods for Efficient Solar Hydrogen Production in Microgravity Environment

Instructor Prep
concepts
Student Protocol
JoVE 杂志
化学
需要订阅 JoVE 才能查看此.  登录或开始免费试用。
JoVE 杂志 化学
Experimental Methods for Efficient Solar Hydrogen Production in Microgravity Environment

1. Preparation of p-InP photoelectrodes

  1. Use single crystal p-InP (orientation (111 A), Zn doping concentration of 5 × 1017 cm-3) as the photoabsorber. For the back contact preparation, evaporate 4 nm Au, 80 nm Zn and 150 nm Au on the backside of the wafer and heat it to 400 °C for 60 s.
  2. Apply Ag paste to attach the ohmic contact to a thin-plated Cu wire. Thread the wire to a glass tube, encapsulate the sample and seal it to the glass tube using black, chemical resistant epoxy.
  3. In order to remove native oxides, etch the 0.5 cm2 polished indium face of p-InP for 30 s in 10 mL of bromine/methanol solution (0.05 % w/v), rinse the surface with ethanol and ultrapure water for 10 s each and dry the sample under nitrogen flux. Prepare the solutions from ultrapure water and analytical grade chemicals with an organic impurity level below 50 ppb.
    CAUTION: Bromine causes acute toxicity upon inhalation, skin corrosion and acute aquatic toxicity. Wear protective equipment such as safety glasses, gloves and lab coat. Work under the fume hood. Methanol is flammable, causes acute toxicity (oral, dermal and inhalation) and is known to cause specific target organ toxicity. Wear protective equipment such as safety glasses, gloves and a lab coat. Work under the fume hood.
  4. Subsequently, condition the p-InP electrode photoelectrochemically in a standard three-electrode potentiostatic arrangement. Use a borosilicate glass cell with a quartz window as a photoelectrochemical cell in order to illuminate the sample with a white-light tungsten halogen lamp (100 mW/cm2) during the procedure.
  5. Adjust the light intensity with a calibrated silicon reference photodiode.
  6. Prepare a 0.5 M HCl solution and purge it in the photoelectrochemical cell with nitrogen of 5.0 purity for 15 min.
  7. Use potentiodynamic cycling between -0.44 V und +0.31 V at a scan rate of 50 mV s-1 for 50 cycles to photoelectrochemically condition the sample under continuous illumination.
    CAUTION: Hydrochloric acid causes serious eye damage, skin corrosion and it is corrosive to metals. Furthermore, it possesses specific target organ toxicity following single exposure. Wear protective equipment such as safety glasses, gloves and a lab coat. Work under the fume hood.

2. Fabrication of rhodium nanostructures

  1. Employ shadow nanosphere lithography (SNL)11,12 for the formation of rhodium nanostructures on the p-InP photoelectrode. In order to create the polystyrene masks on the p-InP electrode, obtain mono-dispersed beads of polystyrene (PS) sized 784 nm at a concentration of 5% (w/v) and dissolve them in ultrapure water.
  2. To obtain the final volume of 600 μL, mix 300 μL of the polystyrene bead dispersion with 300 μL of ethanol containing 1% (w/v) styrene and 0.1% sulphuric acid (v/v).
  3. Apply the solution onto the water surface using a Pasteur pipette with a curved tip. In order to increase the area of the monocrystalline structures, turn the Petri dish gently. Carefully distribute the solution to cover 50% of the air-water interface with a hcp monolayer. Leave place for stress relaxation and avoid forming cracks in the lattice during the next preparation steps.
  4. Protect the Cu wire of the photoelectrochemically conditioned p-InP electrodes with parafilm. Place them delicately under the floating closed packed PS sphere mask by carefully taping them to a microscope slide, preventing the samples from rotating. Gently remove residual water with a pipette and by evaporation, causing the mask to be subsequently deposited onto the electrode surface.
  5. Take the electrode out of the Petri dish and gently dry the surface with N2. Store the electrode under nitrogen until rhodium photoelectrodeposition (e.g., in a desiccator).
    NOTE: The protocol can be paused here for up to one week.

3. Photoelectrodeposition of rhodium nanoparticles

  1. For the photoelectrochemical deposit of rhodium nanoparticles through the PS sphere mask, place the electrode in an electrolyte solution containing 5 mM RhCl3, 0.5 M NaCl and 0.5% (v/v) 2-propanol and apply a constant potential of Vdep = +0.01 V for 5 s under simultaneous illumination with a W-I lamp (100 mW/cm2). Electrochemical specifications such as the electrochemical cell, reference and counter electrode are the same as for the photoelectrochemical conditioning procedure.
  2. Rinse the photoelectrode with ultrapure water and dry it under a gentle flow of N2.
  3. In order to remove the PS-spheres from the electrode surface, place the electrodes for 20 min under gentle stirring in a beaker with 10 mL of toluene (the electrode should be covered with toluene). Subsequently, rinse the electrode with acetone and ethanol for 20 s each.
  4. Remove residual carbon from the surface by O2-plasma cleaning for 6 min at a process pressure of 0.16 mbar, 65 W and gas inflows of O2 and Ar of 2 sccm and 1 sccm, respectively.
  5. Prepare the samples up to one week prior to tests in the drop tower and store them until the experiments under N2 atmosphere in the dark (e.g., in a glove bag or desiccator).
    NOTE: The protocol can be paused here for about 1-2 weeks.

4. Photoelectrochemical experiments in microgravity

  1. For the experiments in microgravity environment, contact one of the major drop tower facilities, (e.g., the Centre of Applied Space Technology and Microgravity (ZARM), Bremen Germany).
    NOTE: By employing a catapult system, 9.3 s of microgravity environment can be generated at ZARM with an approached minimum g-level of about 10-6 m·s-2 13. A hydraulically controlled pneumatic piston-cylinder system is used to launch the drop capsule (Figure 1A) upwards from the bottom of the tower. The capsule is decelerated again in a container which is placed onto the cylinder system during the time of free fall.
  2. Use a two-compartment photoelectrochemical cell (filling volume of each cell: 250 mL) for the photoelectrochemical experiments in order to carry out two experiments in microgravity environment in parallel. The front of each cell should consist of an optical quartz glass window (diameter: 16 mm) for illuminating the working electrode (see Figure 1B).
  3. Employ a three-electrode arrangement in each cell for the photoelectrochemical measurements with a Pt counter electrode and an Ag/AgCl (3 M KCl) reference electrode in HClO4 (1 M). Add 1% (v/v) isopropanol to the electrolyte in order to reduce the surface tension and enhance gas bubble release. Use a W-I white light source for illuminating each cell compartment through the optical windows.
    CAUTION: Concentrated perchloric acid is a strong oxidizer. Organic, metallic and non-organic salts formed from oxidation are shock sensitive and pose a great fire and explosion hazard. Wear safety glasses, gloves and a protective lab coat. Work under the fume hood and minimize bench top storage time.
  4. For gas bubble investigations, attach two cameras to each cell via optical mirrors and beamsplitters (e.g., a color camera at the front and a monochromatic camera at the side, see Figure 1) to record gas bubble evolution during free fall of the experiment. For each drop, store the recorded data on an integrated board computer in the drop capsule. Record single pictures at a frame rate of (e.g., 25 fps (color camera) and 60 fps (monochromatic camera)).
  5. The drop capsule is equipped with several boards (Figure 1). Mount the photoelectrochemical set-up and the cameras onto an optical board and attach it to one of the middle boards in the capsule. Use the remaining boards for installment of additional equipment such as potentiostats, light sources, shutter controls and the board computer. Attach a battery supply at the bottom board of the capsule to power the set-up during free fall (Figure 1).
  6. Write an automated drop sequence for the experimental steps which should be controlled and carried out in microgravity environment. The program should be started prior to each drop. Upon reaching microgravity environment, the sequence should automatically start cameras, illumination sources and the electrochemical experiment for the duration of 9.3 s while simultaneously immersing the working electrode into the electrolyte using a pneumatic system (see Figure 1, Table 1).
  7. Investigation of light-assisted hydrogen production on the samples in photoelectrochemical measurements (e.g., cyclic voltammetry and chronoamperometry).
    1. Control the electrochemical parameters by the two potentiostats in the capsule. For optimal resolutions in J – V measurements, use scan rates (dE/dt) of 218 mV/s to 235 mV/s in order to run 3 scan cycles in cycling voltammetry experiments, using voltage ranges of +0.25 V to -0.3 V vs Ag/AgCl (3 M KCl). Employ the initial potential, Ei = +0.2 V vs Ag/AgCl (3 M KCl) and the finishing potential, Ef = +0.2 V vs Ag/AgCl (3 M KCl). To compare the recorded J – V measurements, take the second scan cycle of each experiment for analysis.
    2. In chronoamperometric measurements, use the time scale of generated microgravity environment, 9.3 s, to record the photocurrent produced by the sample. Apply potential ranges of -0.3 V to -0.6 V vs Ag/AgCl (3 M KCl) to compare produced photocurrents.
  8. At the end of each drop, when the drop capsule is decelerated again to zero velocity, use the drop sequence to let the sample be removed from the electrolyte and cameras, potentiostats and illumination sources be switched off.
  9. After retrieving the capsule from the deceleration container, remove the capsule protection shield. Remove the samples from the pneumatic stative, rinse them with ultrapure water and dry them under gentle nitrogen flux. Store them under N2 atmosphere until optical and spectroscopic investigations are carried out.
  10. Exchange the electrolyte in the two cells, ensure the function of all instruments before equipping the cells with new samples and prepare the capsule for another drop experiment.

Experimental Methods for Efficient Solar Hydrogen Production in Microgravity Environment

Learning Objectives

Etching the p-InP surface in Br2/ methanol for 30 s with consecutive photoelectrochemical conditioning of the sample by cycling polarization in HCl is well established in the literature and discussed (e.g., by Schulte & Lewerenz (2001)14,15). The etching procedure removes the native oxide remaining on the surface (Figure 2) and electrochemical cycling in HCl causes furthermore a considerable increase in the fill factor of the cell performance, accompanied by a flat band shift of the p-InP from +0.56 V to +0.69 V15. Furthermore, the passive layer formation during cyclic polarization in HCl photocathodically protects the InP surface from anodic corrosion. After the conditioning procedure, the self-assembly of 784 nm polystyrene latex nanospheres on the p-InP surface is employed for the formation of a colloidal particle monolayer which further on serves as a lithographic mask during the Rh deposition process (Figure 3A,B). Figure 3B shows an AFM image of the photoelectrode after removal of the PS spheres. The application of SNL results in a nanosized, two-dimensional periodic Rh structure with a homogenous array of holes in the metallic, transparent Rh film. The high-resolution AFM image (Figure 3C) illustrates the hexagonal unit cell structure with recognizable grains of Rh. Cross-section profiles in Figure 3D show that the rhodium mesh is homogenously distributed on the p-InP surface with a height of about 10 nm, forming a catalytic layer. Combined high-resolution TEM and FFT analysis were used to determine the lattice plane spacing, the distance of the diffraction points representing the reciprocal lattice space. Our calculations show that the lattice plane is in the order of 2.17 to 2.18 Å, validating the (111) cubic structure of the deposited rhodium (Figure 4). X-ray photoelectron spectroscopy reveals that the nanostructured p-InP-Rh electrode contains an InOx/POx layer, with evidence being provided by the larger InP signal at 128.4 eV. This is not surprising due to the open InP areas resulting from the removal of the PS spheres; here, the InP is directly exposed to the environment (i.e., air and the electrolyte (Figure 5)).

The microgravity environment has been shown to have a significant impact on the electrolysis of water which has been known since the 1960s and the effect of reduced gravity on the motion of bubbles and drops is well-documented (see e.g., reference 16). Studies have been carried out especially within the frame of developing a life support system for space travel which includes a water-electrolyzing component.

Hitherto investigations of water electrolysis under microgravity environment in 'dark' experiments resulted in a stable gas bubble froth layer formation in proximity to the electrode surfaces and the accompanying ohmic resistance increased linearly to the froth layer thickness in both acidic and alkaline electrolytes17,18,19. Additionally, the diameter of the gas bubbles increased and the bubbles adhered to the membrane separating the two half-cells20,21. Furthermore, it was demonstrated that bubble-induced microconvection dominates the mass transfer in microgravity environments8,21 and it has been suggested that mass transfer of the substrate water to the electrode surface controls the process of water electrolysis, which is controlled under normal gravity conditions by the electrode reaction22.

The here employed nanostructured p-InP-Rh photoelectrodes manufactured via SNL could overcome this problem: the photocurrent-voltage measurements do not show significant differences between terrestrially tested samples in 1 M HClO4 and samples tested in 9.3 s of microgravity environment at the Bremen Drop Tower (Figure 6A,B)6. J-V characteristics (Figure 6A), additionally, chronoamperometric measurements (Figure 6B) of the nanostructured samples are nearly identical in terrestrial and microgravity environment. The difference in the open circuit potential (VOC) is attributed to performance differences of the photoelectrodes as shown earlier6. The introduced rhodium catalytic 'hot spots' on the p-InP surface by SNL allow the formation of gas bubbles to occur at distinct spots on the photoelectrode surface, preventing bubble coalescence and increasing the yield of gas bubble release. The addition of 1% (v/v) isopropanol to the electrolyte decreases the surface tension of the electrolyte furthermore, also leading toward favored gas bubble detachment from the electrode surface.

Figure 1
Figure 1: Scheme of the experimental set-up of the electrochemical experiments in microgravity environment. The images show the equipped drop capsule (A) and details of the photoelectrochemical set-up on the second platform of the drop capsule (B). The capsule contains batteries for power supply during free fall (platform 5), the capsule control system for experimental control (platform 4), two W-I light sources and a board computer (platform 3, see Table of Materials), the photoelectrochemical setup including four digital cameras (platform 2), and two potentiostats and two shutter control boxes (platform 1). The four digital cameras in the photoelectrochemical setup (platform 2) allows recording gas bubble formation on the photoelectrode from the front of each electrochemical cell through beam splitters and from the side through mirrors. The photoelectrodes were illuminated through the beam splitters in front of the cell. Via a pneumatic lifting ramp, the photoelectrodes are immersed in the electrolyte immediately before reaching microgravity conditions. This figure has been modified from Brinkert et al. (2018)6. Please click here to view a larger version of this figure.

Figure 2
Figure 2: Tapping-mode AFM topography images of the p-InP surface before and after the surface modifications steps. Panel A shows the p-InP surface before modification procedures, (B) after etching the surface in bromine/methanol solution and (C) after conditioning the sample in HCl. (D) Histogram analysis of the height distribution of terraces on the p-InP sample (blue line), after etching in bromine/ methanol (yellow line) and after conditioning in HCl (red line). Please click here to view a larger version of this figure.

Figure 3
Figure 3: Tapping-mode AFM topography images (see Table of Materials) of the p-InP surface after application of shadow nanosphere lithography. (A) The deposited polystyrene particle monolayer on the p-InP substrate. (B,C) The surface after deposition of rhodium and removal of the polystyrene particles at two magnifications. (D) A height profile of three different spots on the electrode surface was generated to allow the further characterization of the deposited Rh mesh. Please click here to view a larger version of this figure.

Figure 4
Figure 4: High-resolution TEM analysis of photoelectrodeposited rhodium grains on the p-InP electrode at different magnifications (see Table of Materials). The 2D-Fourier transformation images show the corresponding diffraction pattern with a lattice plane spacing of 2.2 Å (111), typical for cubic structures. Please click here to view a larger version of this figure.

Figure 5
Figure 5: X-ray photoelectron spectra of the nanostructured p-InP-Rh photoelectrodes. (A) In 3d core levels; (B) P 2p core levels and (C) Rh 3d core levels. The color coding under the lines refers to the respective composition as illustrated in the legend. This figure has been modified from Brinkert et al. (2018)6. Please click here to view a larger version of this figure.

Figure 6
Figure 6: Results of the photoelectrochemical experiments in microgravity environment. (A) J-V measurements of nanostructured p-InP-Rh photoelectrodes in terrestrial (1 g, red) and microgravity environments (10-6 g, blue) in 1 M HClO4 with the addition of 1% (v/v) isopropanol to the electrolyte at 70 mW/cm2 illumination with a W-I lamp. Differences in the VOC of the nanostructured samples in terrestrial and microgravity conditions are subject to performance differences of the photoelectrodes as shown earlier6. (B) Chronoamperometric measurements of the nanostructured p-InP-Rh photoelectrodes in terrestrial (red) and microgravity environment (10-6 g, blue) in 1 M HClO4 with the addition of 1% (v/v) isopropanol to the electrolyte at 70 mW/cm2 illumination with a W-I lamp. The applied potential was set to -0.09 V vs RHE. The increased signal-to-noise ratio at the end of the measurements is due to the deceleration of the drop capsule after 9.3 s. Please click here to view a larger version of this figure.

Table 1: Detailed experimental sequence for photoelectrochemical hydrogen production hydrogen production in microgravity environment at the Bremen Drop Tower. This table has been modified from Brinkert et al. (2018)6. Please click here to download this file.

List of Materials

12.7 mm XZ Dovetail Translation Stage with Baseplate, M4 Taps (4 x) Thorlabs DT12XZ/M
Beam splitters (2 x) Thorlabs CM1-BS013 50:50 400-700nm
Beamsplitters (2 x) Thorlabs CM1-BS014 50:50 700-1100nm
Ohmic back contact: 4 nm Au, 80 nm Zn, 150 nm Au Out e.V., Berlin, Germany https://www.out-ev.de/english/index.html Company provides custom made ohmic back contacts
Glass tube, ca. 10 cm, inner diameter about 4 mm E.g., Gaßner Glasstechnik Custom made
p-InP wafers, orientation 111A, Zn doping concentration: 5 x 10^17 cm^-3 AXT Inc. Geo Semiconductor Ltd. Switzerland Custom made
Photoelectrochemical cell for terrestrial experiments E.g., glass/ materials workshop Custom made
Matrox 4Sight GPm (board computer) Matrox imaging Ivy Bridge, 7 x Cable Ace power I/O HRS 6p, open 10m, Power Adapter for Matrox 4sight GPm, Samsung 850 Pro 2,5" 1 TB, Solid State Drive in exchange for the 250Gb hard drive
2-propanol Sigma Aldrich I9516-500ML
35mm Kowa LM35HC 1" Sensor F1.4 C-mount (2 x) Basler AG
Acetone Sigma Aldrich 650501-1L
Ag/AgCl (3 M KCl) reference electrode WPI DRIREF-5
Aluminium breadboard, 450 mm x 450 mm x 12.7mm, M6 Taps (2 x) Thorlabs MB4545/M
Beaker, 100 mL VWR 10754-948
Black epoxy Electrolube ER2162
Bromine Sigma Aldrich 1.01945 EMD Millipore
Colour camera (2 x) Basler AG acA2040-25gc
Conductive silver epoxy MG Chemicals 8331-14G
Copper wire E.g., Sigma Aldrich 349224-150CM
Ethanol Sigma Aldrich 459844-500ML
Falcon tubes, 15 mL VWR 62406-200
Glove bags Sigma Aldrich Z530212
Hydrochloric acid (1 M) Sigma Aldrich H9892
Magnetic stirrer VWR 97042-626
Methanol Sigma Aldrich 34860-100ML-R
Microscope slides VWR 82003-414
MilliQ water
NIR camera (2 x) Basler AG acA1300-60gm
Nitrogen, grade 5N Airgas NI UHP300
Ø 1" Stackable Lens Tubes (6 x) Thorlabs SM1L03
O2 Plasma Facility
OEM Flange to SM Thread Adapters (4 x) Thorlabs SM1F2
Parafilm VWR 52858-000
Pasteur pipette VWR 14672-380
Perchloric acid (1 M) Sigma Aldrich 311421-50ML
Petri dish VWR 75845-546
Photoelectrochemical cell for microgravity experiments E.g., glass/ materials workshop
Polystyrene particles, 784 nm, 5 % (w/v) Microparticles GmbH 0.1-0.99 µm size (50 mg/ml): 10 ml, 15 ml, 50 ml
Potentiostats (2 x) Biologic SP-200/300
Pt counter electrode ALS-Japan 12961
Rhodium (III) chlorid Sigma Aldrich 520772-1G
Shutter control system (2 x)
Silicon reference photodiode Thorlabs FDS1010
Sodium chlorid Sigma Aldrich 567440-500GM
Stands and rods to fix the cameras VWR
Sulphuric acid (0.5 M) Sigma Aldrich 339741-100ML
Telecentric High Resolution Type WD110 series Type MML1-HR110 Basler AG
Toluene Sigma Aldrich 244511-100ML
Various spare beakers and containers for leftover perchloric acid etc for the drop tower VWR
W-I lamp with light guides (2 x) Edmund Optics Dolan-Jenner MI-150 Fiber Optic Illuminator
CM-12 electron microscope with a twin objective lens, CCD camera (Gatan) system and an energy dispersive spectroscopy of X- rays (EDS) system) Philips
Dimension Icon AFM, rotated symmetric ScanAsyst-Air tips (silicon nitride), nominal tip radius of 2 nm Bruker

Lab Prep

Long-term space flights and cis-lunar research platforms require a sustainable and light life-support hardware which can be reliably employed outside the Earth’s atmosphere. So-called ‘solar fuel’ devices, currently developed for terrestrial applications in the quest for realizing a sustainable energy economy on Earth, provide promising alternative systems to existing air-revitalization units employed on the International Space Station (ISS) through photoelectrochemical water-splitting and hydrogen production. One obstacle for water (photo-) electrolysis in reduced gravity environments is the absence of buoyancy and the consequential, hindered gas bubble release from the electrode surface. This causes the formation of gas bubble froth layers in proximity to the electrode surface, leading to an increase in ohmic resistance and cell-efficiency loss due to reduced mass transfer of substrates and products to and from the electrode. Recently, we have demonstrated efficient solar hydrogen production in microgravity environment, using an integrated semiconductor-electrocatalyst system with p-type indium phosphide as the light-absorber and a rhodium electrocatalyst. By nanostructuring the electrocatalyst using shadow nanosphere lithography and thereby creating catalytic ‘hot spots’ on the photoelectrode surface, we could overcome gas bubble coalescence and mass transfer limitations and demonstrated efficient hydrogen production at high current densities in reduced gravitation. Here, the experimental details are described for the preparations of these nanostructured devices and further on, the procedure for their testing in microgravity environment, realized at the Bremen Drop Tower during 9.3 s of free fall.

Long-term space flights and cis-lunar research platforms require a sustainable and light life-support hardware which can be reliably employed outside the Earth’s atmosphere. So-called ‘solar fuel’ devices, currently developed for terrestrial applications in the quest for realizing a sustainable energy economy on Earth, provide promising alternative systems to existing air-revitalization units employed on the International Space Station (ISS) through photoelectrochemical water-splitting and hydrogen production. One obstacle for water (photo-) electrolysis in reduced gravity environments is the absence of buoyancy and the consequential, hindered gas bubble release from the electrode surface. This causes the formation of gas bubble froth layers in proximity to the electrode surface, leading to an increase in ohmic resistance and cell-efficiency loss due to reduced mass transfer of substrates and products to and from the electrode. Recently, we have demonstrated efficient solar hydrogen production in microgravity environment, using an integrated semiconductor-electrocatalyst system with p-type indium phosphide as the light-absorber and a rhodium electrocatalyst. By nanostructuring the electrocatalyst using shadow nanosphere lithography and thereby creating catalytic ‘hot spots’ on the photoelectrode surface, we could overcome gas bubble coalescence and mass transfer limitations and demonstrated efficient hydrogen production at high current densities in reduced gravitation. Here, the experimental details are described for the preparations of these nanostructured devices and further on, the procedure for their testing in microgravity environment, realized at the Bremen Drop Tower during 9.3 s of free fall.

Procedure

Long-term space flights and cis-lunar research platforms require a sustainable and light life-support hardware which can be reliably employed outside the Earth’s atmosphere. So-called ‘solar fuel’ devices, currently developed for terrestrial applications in the quest for realizing a sustainable energy economy on Earth, provide promising alternative systems to existing air-revitalization units employed on the International Space Station (ISS) through photoelectrochemical water-splitting and hydrogen production. One obstacle for water (photo-) electrolysis in reduced gravity environments is the absence of buoyancy and the consequential, hindered gas bubble release from the electrode surface. This causes the formation of gas bubble froth layers in proximity to the electrode surface, leading to an increase in ohmic resistance and cell-efficiency loss due to reduced mass transfer of substrates and products to and from the electrode. Recently, we have demonstrated efficient solar hydrogen production in microgravity environment, using an integrated semiconductor-electrocatalyst system with p-type indium phosphide as the light-absorber and a rhodium electrocatalyst. By nanostructuring the electrocatalyst using shadow nanosphere lithography and thereby creating catalytic ‘hot spots’ on the photoelectrode surface, we could overcome gas bubble coalescence and mass transfer limitations and demonstrated efficient hydrogen production at high current densities in reduced gravitation. Here, the experimental details are described for the preparations of these nanostructured devices and further on, the procedure for their testing in microgravity environment, realized at the Bremen Drop Tower during 9.3 s of free fall.

Tags