Sunteți pe pagina 1din 40

3D Graphene Foam for Energy and Environmental Applications

Three dimensionally free-formable graphene foam with a large surface area (994.2 m2/g),

excellent conductivity (2.39 S/cm), reliable mechanical properties (E = 239.7 kPa) and tunable

surface chemistry. The 3D graphene foam has been evaluated towards electrical, energy and

environmental related issues.

Abstract: Three-dimensional assemblies of graphene have been considered as promising starting

materials for many engineering, energy and environmental applications due to its desirable

mechanical properties, high specific area, superior thermal and electrical transfer ability.

However, little has been done to introduce designed shapes into scalable graphene assemblies. In

this work, we show here a combination of conventional graphene growing technique – chemical

vapor deposition with additive manufacturing. Such novel synthesis collaboration enables a

hierarchically constructed porous 3D graphene foam with large surface area (994.2 m2/g),

excellent conductivity (2.39 S/cm), reliable mechanical properties (E = 239.7 kPa) and tunable

surface chemistry that can be used as strain sensor, catalyst support, solar steam generator.
1. Introduction

The introduction of graphene since it was first prepared in 2004 heralded a new epoch in

research. The fascinating properties such as large surface area, high thermal conductivity, rather

enhanced electrical conductivity mediated applications in many engineering, energy, and

environmental related issues. [1] Incorporating binder into fabrication of 3D architecture may

significantly block the active surface area and causing degradation of electrical behaviour. To

achieve the promise of using bare graphene, processing that preserved its unique properties is

demanded. Flat 2D graphene sheets tend to agglomerate due to their strong π-π interlayer

reactions, which prevents the realization of high specific area and fast carrier mobility.[2]

Agglomeration should therefore be avoided, such as by combining individual graphene sheets

into an object of macroscopic scale. Moreover, the exceptional mechanical properties of 2D

graphene translates to a robust 3D graphene assembly.[3] Current 3D graphene assemblies,

fabricated by solution-based gelation of graphene oxide sheets followed by chemical reduction

process or pyrolysis routes (template-free or template assisted), are restricted to laboratory

scale.[4] Moreover, for reassembly of exfoliated graphene oxide sheets, pre-set laminar order

and inferior quality of end products restricts it’s practical applications to a certain degree.[5]

Chemical vapor deposition, among various synthesis methods, has been regarded as the most

reliable approach to grow highly crystalline 2D graphene film and to directly form covalently

bonded 3D framework onto templates.[6] It was introduced to the public first in 2009 by three
independent groups showing the ability of producing single- or few-layer of graphene with

centimeter scale.[7] Such graphene monoliths grown from macroscopic templates later emerged

as superior multi-functional supports. Recent literature explored the potential of such materials in

the fields ranging from wearable electronics,[8] energy storage devices,[9] electrochemistry

framework,[3b, 10] to the sensors,[11] thermal management,[12] and other applications.[13]

The choice of sacrificial template determines the interconnected graphene network and

macroscopic architecture. Existing templates ranges from metallic foam/foil/powder products to

various salts and oxides in the forms of powders and aerogels.[14] [15] However, for foam/foil

and powder templates having relatively large pore size in the range of hundreds of micrometers,

as-fabricated graphene monoliths are not given enough bonding density unless enhanced by a

polymer matrix. Efforts on fabricating templates with a finer interconnected network brought

nano-porous templates (seashell,[16] aerogel,[15e] etc.) to the stage. Attempts combining such

bulk templates with nano-porosity and CVD route did produce lightweight, freestanding 3D

graphene with ultra-high stiffness. However, classical acid-catalyzed sol-gel process generally

produce a pre-determined microstructure. Also, aerogels like silica aerogels usually contain lots

of unreacted hydrophilic silanol (Si-OH) groups, causing volume change. Issues of shrinkage and

cracking leads to difficulties in the use of large volume or high aspect ratio templates. Despite

advances made in the previous documented work; several challenges still remained unsolved.
The gradient distribution of precursor concentration and temperature along the gas flow and

heating source significant influenced the uniformity and quality of graphene grown on the

substrate, thus hampering the scalability of end products and yields. In addition, it also has the

drawback of a lack of complexity in macroscopic design and microscopic porosity optimization

to increase the exposed surface area. Accordingly, the production capacity as well as a tedious

removing process can further increase the cost.[17] To address this problem, additive

manufacturing is a promising alternative to not only aim for the industrial production of

graphene monolith, but also demonstrate the possibility of customized design of graphene-based

electronics, devices, filter membrane and beyond. Previous works regarding 3d printing of

carbon-based materials have been well-documented using direct ink writing technique due to its

wide range of materials selection.[18] Digital light processing (DLP) can print out structures

with a smooth surface finish, and higher resolution (tens of micrometers), yet it has not been

explored for fabricating high quality porous template for 3D graphene foam. In DLP, the acylate-

based resin is mixed with filler materials and can be polymerized via photopolymerization layer

by layer. On the other hand, SiO2 (Silica) powder and its aerogel are an appropriate template for

graphene nucleation and growth as illustrated by Zhi et. al. and Huang et. al.. The use of DLP for

porous silica ceramic have not received much attention. As compared to fused silica or quartz

and glass,[20] porous silica can be effectively removed under mild condition. Considering

porous silica as a sacrificial template can also significantly reduce processing time and
complication. Therefore, we report here a facile synthesis route via 3d-printed silica sacrificial

template for a bi-continuous porous 3D graphene foam with high surface area, excellent

conductivity, superior mechanical properties and customizable design towards various

applications. In order to obtain high-quality porous silica template, optimal solid loading and

appropriate rheology behaviour were evaluated. To be effective, compatibility of silica filler with

the polymer resin was tested to ensure good layer adhesion, reasonable curing time and stability

during long-term printing. The optimal debinding process leading to a rational control over

porosity was determined after sintering study. A class of complex structures that cannot be

fabricated by conventional methods was demonstrated. 3D graphene foam grown on such

template followed the macroscopic design and was proven to facilitate better dissolution as

compared to many state-of-the-art monolith templates. Four separate applications towards

electrical (strain-induced resistance change), energy (overall water splitting) and environmental

(sea water desalination/steam generation and oil adsorption) related issues were investigated to

demonstrate the wide-ranging possibilities of our work.


Figure 1. Design consideration of templates: (a) Macroscopic porous foam template; (b) Solid template with
microscopic nano-pores; (c) template with stratified pores; (d) Preparation methods involved in fabricating 3D
graphene assembly. (I) DLP printing of silica green body (Macroscopic structure design) and its Scanning Electron
Microscopic(SEM) image; (II) Debinding of polymer additives (or ligand removal) and sintering procedure for
porosity formation (Microscopic porosity formation) with its Micro-CT scanning image; (III-IV) Chemical vapor
deposition and wet etching of sacrificial silica template to form 3D graphene assemblies with its SEM image
(Bottom row III) and TEM image (Bottom row IIV),
Figure 2. (a-c) optical images of (a)printed, (b)sintered, (c)CVD grown silica (d) 3D freestanding graphene; (e-g)

SEM images of (e) as-printed silica, (f) sintered silica, (g) CVD grown, (h) 3D freestanding graphene gyroid slab; (i)

micro-CT scanning of CVD grown silica to show the cross-sectional structures.

2. Results and discussion

2.1. Graphene monolith


The fabrication process for 3D graphene foam was summarized in Figure 1, containing 3D

printing of sacrificial silica template and the growth of high-quality graphene. Porous silica

templates with complex designed shapes were first prepared by digital light process method with

photo polymerization of UV-curable resin containing ceramic particles in suspension. The resin

must be highly loaded to produce reliable ceramic parts. However, it must have a low viscosity

(< 5000 cps or < 5 Pa.s) for successful recoating and self-levelling following the requirements of

many commercial products. Therefore, prior to the printing, the rheology behaviour and resin

stability were studied. Resin stability is important to ensure good interlayer adhesion, resulting in
better load transfer from the resin to the filler, hence giving superior strengthening effects. In

Figure S1, the viscosity at shear rates of 0.1 and 180 s-1 showed minor changes over 50 cycles

within 2 days, indicating excellent stability. High loading of solid silica particles did not affect

the resin flowability much, as illustrated by the low viscosity value (~1 Pa.s). Solid loading of

the resin was determined by TGA analysis to be 67.8 wt% silica as shown in Figure S2. Different

computer-aided design files were then sent to the printer, sliced and printed in a layer by layer

manner. For efficient printing, layer thickness was set to be 100 μm. After obtaining the printed

structures, these objects were subjected to further debinding and sintering processes. From TGA

analysis, polymer decomposition started slowly at 100 ℃ and rapidly from 400 ℃ to 500 ℃.

Therefore, extra dwelling steps at temperature 200, 400 and 500 ℃ were added to ensure

complete polymer burnout. A rational polymer burn-off step can be therefore established

according to this result. To study sintering conditions, the as-prepared structures were then

sintered under atmospheric conditions at various temperatures (1250, 1350 and 1450 ℃) and

soaking times (5, 10 hours). Comparisons between different thermal hydrolysis conditions can be

found in Figure S3. A slow ramping rate at 1 ℃/min was selected to maintain uniform pore

distribution.[21] The formation of 3D graphene required the CVD method. CVD was performed

using CH4 at a flow rate of 10 sccm as the precursor and elevate to 1100 ℃ for 1 hour.[15e]

Changing the flow rate to 20 sccm or increasing the time range to 2 hours will distort or block

the original features respectively (shown as Figure S4). A gyroid slab structure (2.7 x 1 x 0.4

cm3) was adopted for the following demonstration unless additional annotation was added as

presented in Figure 2. The observed layers stacking the vertical direction illustrated that the

adjacent two layers were cured and fused cleanly after printing and thermal hydrolysis. On top of

this, Figure 2i demonstrated a graphene coated silica gyroid cube with no internal cracks via
micro-CT cross-section scans. After CVD, silica was etched away with aqueous HF solution,

obtaining the freestanding graphene foam. Fabrication of crack-free and phase-pure bi-

continuous complex graphene foam cannot be obtained from conventional ceramic shaping

methods as shown in Figure S5b,5d. Figure S5a showed the scanning electron microscopy (SEM)

of the colloidal silica resin used in this study. Sintering of the spin-coated resin film results in

completely different cracking behaviour as compared with 3D printed fine structure with same

feature size as demonstrated in the same figure. Limited structures can be provided that is not

prone to sintering cracks via conventional casting. Additive manufacturing instead rationally

applies shape design to enable a crack-free sintered body. Moreover, as for removal of the

sacrificial template, the 3D graphene assembly from conventional casting leads to the anisotropic

etching behaviour of the coated silica template that crack severely after wet etching step.

Effective etching took place at a few hundred micrometers from the contact surface as indicated

in the elemental line-scan analysis. The same process conducted on a 3D-printed template with

finer feature size led to a uniform and mild reaction procedure that instead resists volume

shrinkage or cracking. The SEM image and EDX mapping of the microstructure of the fully

processed 3D graphene foam showed the remarkably homogeneous grain size and shape, as

determined by the nanoparticle building blocks in Figure S5f-g.

As shown in Figure 3, final products showed an interconnected network of foam-like graphene

and density of about 18 mg/cm3. Figure 3a-b suggested that a similar microstructure and pore

distribution can be maintained after chemical etching of sacrificial template. The microscopic

features like step edges were still preserved (Figure 2h). The foam-like graphene network is built

up by few-layered/multi-layered graphene (Figure 3c-f). EDX elemental mapping was conducted

on the surface of 3D graphene foam. From the results, elemental C fully covered the original
microstructures (Figure 3b). Elemental quantification exhibited 100 % C, highlighting the

excellent purity of as-fabricated sample. Similar extrapolation echoed well in the XPS full

pattern spectra where only 1.27 at% oxygen impurity can be detected. High resolution analysis of

O 1s region suggested the surface adsorbent (-OH, 533.6 eV) mainly contributed to such minimal

impurity (Figure 6i) .[22] To further investigate the quality of the graphene, Raman spectroscopy

was presented as shown in Figure 3f. A characteristic 2D band peak shift was detected in the

Raman spectra as a prove of the existence of few layer/multilayered graphene.[15e] Comparison

between 3D graphene foam and other carbon-based materials in the intensity ratio of the D band

and G band further highlights the formation of high-quality graphene. The position of the 2D

bands located as around 2690 cm-1 as well as the I2D/IG ratio (about 0.5) indicated that few

layer/multilayered graphene should be formed. Such inference matched well with high-resolution

XPS C 1s spectra (Figure 6g). A large specific surface area of 994.2 m2/g was determined by the

BET analysis. A well-matched pore distribution in the silica template with the as-processed

graphene foam was observed, showing the co-existence of mesopores and micropores. The

porosity of 3D printed graphene foam was 99.2 %, by considering the density of graphite (which

is 2.23 g/cm3). The highly porous graphene foam can potentially be of great value to energy

storage and mass transport related applications. From a practical point of view, the structural

robustness of the synthesized porous graphene foam is critical. The interconnected graphene

network was tested to be also robust and durable under simulated working conditions. Young’s

modulus, given by the slope of the tangent of the stress-strain curve, is about 239.7 kPa at ρ = 18

mg/cm3 for as-fabricated 3D assembly and stands out from many other carbon-based monolith

(Figure S6 and Figure 4c). The stress remained almost constant at 55% compression strain for

different loop, suggesting elasticity and flexibility (Figure 4b). Moreover, the strength-density
chart (Figure S7) indicated such foam possess better strength-to-weight ratio than flexible

polymer foams and is comparable to rigid polymer foam. The conductivity of 3D graphene foam

is determined as 2.39 S cm-1 at room temperature by four-probe measurements, showing a

superior conductivity as compared to other carbon-based materials at same density level (Figure

4d). A well-constructed bi-continuous network and low impurity level may contribute to the

excellent electrical behaviour.

Figure 3. (a) SEM image of 3D freestanding graphene foam; (b) corresponding EDX image; (c) TEM images of 3D

freestanding graphene foam; (d) XPS summary spectra of 3D freestanding graphene foam; (e) XRD pattern of CVD

grown silica (denoted as Graphene @ SiO2) and 3D graphene foam; (f) comparison of Raman spectra between

carbon-based materials and 3D freestanding graphene foam.


2.1.1. Strain-induced resistance change

To demonstrate the possibilities of this freestanding graphene foam, several well-known

property tests were herein presented. First, foam-like graphene structures with flexibility and

robustness under mechanical strain and stress can be crucial in application of sensors, wearable

devices and actuators. When subjected to compressive strain, it should function as an electrical

resistor with a drop in resistance. Figure 4e-f demonstrated a set-up of compression-assisted

resistivity testing and the resistance change with strain using a gyroid slab graphene. Bending

induced signals were collected based on a determined value of 32 ° during the downward

movement. The resistance variation (~ 9 %) was observed to be minimal during the repeating

cycles. Under another compression strain-induced test, the sensitivity was first increase

significantly (DR/R0 ª 90 % at a strain of 25 %) and showed less sensitivity to higher

compressive strain. The starting resistance (~300 ohm) can be higher than that of original 3D

graphene foam (measured by bulk plate) due to the contact resistance and higher porosity of the

gyroid design. The results again confirmed a formation of durable and robust graphene foam

with no degradation of its characteristic properties.


Figure 4. (a) BET surface area N2 adsorption-desorption isotherm plots measured over silica template and 3D

graphene foam; (b) the compressive stress at 10 %, 20 %, 40 % and compressive strain plotted versus cycle number;

(c) ρ versus E of the first cycle of 3D graphene foam (this work), GO aerogel,[18b] GO+PMA+PEG,[18a] Zinc-

assisted graphene monolith,[5b] Aerographite,[23] and GO+CNT;[24] (d) Electrical conductivities of 3D graphene

foam (this work), 3D-printed GO,[25] Zinc-assisted graphene monolith,[5b] Aerographite,[23] and CNT aerogel;[26]

(e) Resistance changes with bending strain exhibiting similar recovery behavior at 32 °bending motion. (Structure:

ribbon; length: 3 cm; width: 0.2 mm; thickness: 500 μm); (f) Resistance changes with compression strain up to 60 %

compression. (Structure: gyroid slab; length: 2.7 cm; width: 1 cm; thickness: 0.4 cm).
Figure 5. (a-f) SEM images of NiFe LDH/CC, NiFe LDH/GP and NiFe LDH/GF; (g-h) TEM images of NiFe LDH;

(i-n) electrochemical performance test of all sample for: (i-k) OER; (l-n) HER.
2.1.2. Catalytic Behaviour

Second, cellular carbon-based monolith has been regarded as the most promising electrode to

support catalyst due to its extraordinary specific surface area (SSA). A robust graphene foam

with excellent conductivity should be explored for wider application provided it inherited high

SSA. As compared to powder-based research regarding the same materials, fabrication of a high-

quality graphene monolith can benefit such applications. Herein, the as-prepared graphene foam

was confirmed to have a surface area as high as 994.2 m2 /g by BET. A large surface area and

hierarchically constructed porous structure can be essential to retain outstanding electrochemical

performance with high mass loading for energy conversion and storage reactions. Recently,

catalyst materials like transition metal hydroxides have long been well studied and proven to

exhibit sound activity towards overall water splitting. Therefore, nickel/iron double layer

hydroxides (NiFe LDH) were grown on a thin 3D graphene gyroid slab (2.7 x 1 x 0.1 cm3,

denoted as NiFe LDH/GF) via a typical urea hydrolysis of a mixture nickel nitrate and iron

nitrate. In a control experiment, other commonly used carbon substrates such as carbon cloth

(denoted as NiFe LDH/CC) and graphite paper (denoted as NiFe LDH/GP) were adopted. The

deposition of NiFe LDH was uniformly coated over the gyroid slab (Figure 5c, 5g) while coating

only took place at the very top layer of carbon cloth (Figure 5a, 5e) and inconsistency existed in

the surface of graphite paper (Figure 5b, 5f). Regarding the loading mass, 17 mg/cm2 can be

obtained which is almost 10 times higher than carbon cloth and graphite paper. Electrochemical

properties of the three substrates and coated substrates were investigated using 1 M KOH in a

three-electrode cell for overall water-splitting behaviour (Figure 5). Oxygen evolution reaction

performance was recorded in Figure 5i-5k. From the polarization curves in Figure 5i, the

graphene foam (392 mV @ 10 mA/cm2) and NiFe LDH/GF (198 mV @ 10 mA/cm2) exhibited
the lowest overpotential values as compared to other substrates selected, indicating large increase

in loading mass did not retard charge transfer efficiency. Cyclic voltammetry (CV) scanning was

also provided to show the electrochemical double-layer capacitance (Cdl). Notably, the

electrochemical active surface area was significantly improved by coating of NiFe LDH on

gyroid graphene foam, changing from 30.7 to 69.9 mF/cm2. The drastic increase surpassed that

of NiFe LDH/CC (1.57 times higher than blank substrate) and NiFe LDH/GP (1.54 times higher

than blank substrate). The Tafel slope was also given in Figure S8 and NiFe LDH/GF displayed

the lowest value of 74 mV/dec. From the OER equation, a good OER electrocatalyst system

should have favourable the electron transfer performance, excellent conductivity and large ECAS

property. All these factors will influence the electron transfer performance. The OER

performance of this catalytic system is highly favourable compared to many reported works.

Hydrogen evolution reaction (HER) was also conducted to get the activity of the same sample

class. The NiFe LDH/ GF yet again provided the best HER performance with an overpotential of

230 mV @ 10 mA/cm2 and extremely high ECSA of 155.6 mF/cm2. All the results were

summarized in Figure 5j-m and Figure S8 for a better comparison. These results supported the

proposal of adopting such graphene foam for various electrochemical energy storage device

where a conductive electrode with high specific/electrochemical active surface areas is desired.
Figure 3. (a-c) solar steam generation of as-fabricated samples: GF and Treated GF; (d) efficiency towards solar

steam generation for all samples; (e) UV-Vis spectra for light absorption of Treated GF; (f) Water contact angle

measurements of GF and Treated GF; (g-i) high resolution XPS spectra of: (g-h) C 1s and (i) O 1s; (j) Absorption

and removal of Hexane: (I) mixture of red-stained Hexane and blue-stained DI water, (II) drop the as- fabricated

graphene foam into the mixture absorption, (III) absorption of Hexane, (IV) taking out the graphene foam, (V)

burning away of Hexane (Insert: after the burning).


2.1.3. Thermal management and surface chemistry

Thirdly, apart from excellent electrical conductivity and high surface area, graphene materials

have their versatile surface chemistry and thermal management ability. Solar energy which is

regarded as the largest source of renewable energy that can be used for daily life. Thermal

management is one of the most direct paths of utilizing solar energy. Solar technologies such as

solar steam generator has been well developed and has great potential in global water supply.

Numerous works have listed carbon-based heat absorbers for sea water desalination or solar

steam generation. The atomically smooth and thin nature of graphene facilitated efficient

transport of water through nanochannels, microscopic porosity and defects. In Figure 6, we

simply demonstrated our 3D gyroid graphene foam (2.7 x 1 x 0.4 cm3) as a floating heat

absorber system (denoted as GF). The gyroid macroscopic structures were also believed to

improve water uptake behaviour. As hydrophilic contact with open porosity can serve well for

water transportation ability, surface treatment (immersing into an aqueous solution of 50 %

HNO3, denoted as Treated GF) was further applied on the graphene surface, modifying the

surface from originally hydrophobic to hydrophilic. The wettability of GF and Treated GF were

recorded by water contact angle measurements. Treated GF showed transformation to a

hydrophilic behaviour while untreated GF exhibited a contact angle of 126.7 ° (Figure 6f).

Changes in surface chemistry originated from the increase of hydrophilic functional group such

as -OH. As indicated in Figure 6i, surface oxygen level increased recognizably, confirmed by the

increase in peak intensity of high-resolution O 1s spectra. The signal assigned to -OH group was

amplified as well. Another observation was the decrease of total C 1s signal intensity and

increase of C-O, C=O and O-C=O (285.6 eV, 286.7 eV and 288.6 eV, respectively) peaks.[22]

The hydrophilic surface channelled the water through the porous structure and favoured the
water desalination process. To monitor the optical adsorption ability, the UV-Vis spectra was

also measured. It showed 98 % adsorption efficiency in the near-infrared, visible and near-

ultraviolet region (Figure 6e). An IR camera was used to record the temperature distribution. The

maximum temperature under one sun illumination can rise to 43 ℃ within a few minutes and

remained constant afterwards for Treated GF, while water under the same illumination only

experienced a slight temperature increase (~ 3 ℃) as indicated in Figure 5c. Such improvement

in thermal adsorption can be attributed to: (a) hierarchically constructed porosity form a heat

localization layer to confine the radiation; (b) a rough surface and surface modification promoted

lower light reflectance and thermal management.[6c] The evaporation rate was plotted according

to the weight loss as a function of time under 1 sun irradiation. The cumulative weight loss is

linear with illumination time, demonstrating a steady evaporation rate as in Figure 6a-b. Largest

weight loss and evaporation rate over the same illumination time period were obtained by

Treated GF, manifesting the best steam generation performance. Evaporation rate of water

facilitated by Treated GF is about 1.436 kg/m2 h, which is 3.55 times higher than that of pure

water. For GF, the evaporation rate showed a lower value of 1.321 kg/m2 h. The solar energy

conversion efficiency therefore can be calculated. While used for solar steam generation, such

floating system displayed an efficiency of 84.5% and 93% for GF and Treated GF, respectively.

As a controlled experiment, pure water was also tested for its conversion efficiency, which

showed an efficiency of 25.5%. The dark field performance was also presented in the support

information together with a comparison with other reported solar steam generation systems.

Similar surface modification strategies can also be adjusted case by case when being used for

other applications like oil-water separation and filtration. Moreover, GF itself as a hydrophobic

system and can also be used for oil/solvent absorption as demonstrated in Figure 6j. The solvent
absorption of red-stained Hexane by the as-fabricated graphene foam was demonstrated (Figure

6j). Such material hence showed extensive applicability and high efficiency for oil/solvent

contaminant adsorption and recyclability.

3. Conclusion

Sacrificial silica template with complex-designed structures prepared by 3D printing techniques

enables the fabrication of crack-free and phase-pure bi-continuous complex graphene foam. As-

fabricated hierarchically porous graphene foam inherited a large surface area (994.2 m2/g),

excellent conductivity (2.39 S/cm), reliable mechanical properties (E = 239.7 kPa) and tunable

surface chemistry. Such 3D graphene foam was demonstrated as strain sensor, catalyst support

and solar steam generator. This work represents a step toward fulfilling the grand promise of 2D

graphene in a 3D world. A major constraint to the utilization of graphene up to now has been the

limited form of the material due to conventional fabrication methods. The fabrication of high-

quality 3D graphene enables greater adoption in wide-ranging fields from wearable sensors,

energy storage, water supply etc. All characterizations and demonstrations shed a light on this

3D graphene that is being further explored.

4. Experimental Section

Materials: Silica powder, photo-initiator and viscosity modifier like Diphenyl (2,4,6 trimethyl

benzoyl) phosphine oxide (TPO), ethoxylate 1,6-hexanediol diacrylate (E-HDDA), and

ethoxylated trimethylolpropane triacrylate (E-TMPTA) were purchased from Sigma Aldrich.

Variquat CC 42 NS was kindly provided by Evonik and served as the dispersant in this
experiment. Sigma Aldrich also provided isopropyl alcohol (IPA) and hydrofluoric acid (HF, 45

wt%) used for washing and etching.

Preparation of DLP Silica Resin: In an amber colour bottle, weigh 53 grams Silica powder, 3 mL

of Variquat CC 42 NS and 25 mL of photocurable resin (consisting of E-HDDA and E-TMPTA

at a ratio of 3:22 with 2 wt% TPO photo initiator). The photocurable slurry is then homogenized

and printed using Asiga Max at 50 to 100 μm layer height setting. E-HDDA and E-TMPTA were

added to reduce viscosity and to improve curing properties.

Stereolithography and Heat Treatment of Silica: For stereolithography, the resin was processed

with Asiga® MAX (commercial printer with UV source of 385 nm). The built structure can have

the dimension up to 119 x 67 x 75 mm3.All the structures presented in this work were either

downloaded from Thingiverse (Available at https://www.thingiverse.com/) or drawn using

Fusion 360. Before printing, the STL file was sliced by the software that comes with the printer.

The layer thickness was set to be 100 μm for each print. Burn-in layer time was 7 seconds and

exposure time for each layer was 1-1.5 seconds. The printed structure was removed gently from

the building platform with transparency film purchased from Suremark. The structure was

directly moved to a 100 ml beaker and immersed in IPA. An ultrasonic bath was applied to

remove any unpolymerized residue and to reveal the designed shape. This step was repeated for

five times before subjected to drying at room temperature. Our preliminary results and

experiences determined the debinding and sintering condition. A debinding condition (soak at

200, 400, 500 oC for 2 hours respectively) was applied before the sintering of silica (Carbolite

AAF 1100 Furnace). After that, the debinded silica structures were later transferred to the high-

temperature furnace (Carbolite HTF 1800 Furnace) and annealed at 1250-1450 oC for 5-10 hours.
Fabrication of Graphene foam: The 2D graphene was grown onto the gyroid silica template

under a gas flow containing 10-20 sccm CH4 for 1-2 hours. The heating and cooling were done

in the same tube under a carrier gas flow of H2 (50sccm) and Ar (300 sccm) using SHW-300C

hot-wall CVD system. After cooling to the room temperature, the whole product was immersed

in 15% aqueous HF solution overnight to remove the silica template. After etching, the HF

solution was drawn out and changed to DI water to wash away the residual solution. Freeze

drying with liquid N2 was applied to obtain the final 3D graphene sample.

NiFe LDH @ 3D Graphene Foam, Carbon Cloth, and Graphite Paper Electrode: Ni1.5Fe0.5

LDHs spheres were grown at various carbon-based substrates using the well-documented

hydrothermal method. All the substrates were cut into 1cm2 piece and put in the hydrothermal

vessel after washing with deionized water and ethanol. For every 40mL aqueous solution added

into the 50 ml Teflon-lined autoclave, 0.1745 g Ni(NO3)2·6H2O, 0.0808 g Fe(NO3)3·9H2O,

0.24 g urea (4 mmol), and 0.05925 g NH4F were included. The hydrothermal process was

maintained at 120 oC for 12 hours and subsequently cooled to room temperature. Afterward, the

coated samples were collected and rinsed with deionized water and ethanol for several times,

followed by drying at fume hood overnight, yielding NiFeLDH @ GF (Gyroid Graphene Foam),

NiFeLDH @ CC (Carbon Cloth), NiFeLDH @ GP (Graphite Paper).

Characterization: Ceramic resin stability measurements were performed on TA Instruments

DHR-2 Rheometer using 40mm diameter parallel plates at 500 μm measurement gap over 48

hours. The apparent viscosity of the suspensions was measured at 25 oC in a shear rate sweep

mode with shear rate ranging from 0.01 – 180 s-1 over 48 hours. The surface morphology of as-

fabricated electrodes was observed using a scanning electron microscope (SEM) with an

acceleration voltage of 10 kV with EDX mapping using the same voltage (Zeiss; FESEM Supra

40). X-ray powder diffraction (XRD) patterns were obtained by Bruker D8 Advanced
diffractometer system with Cu Kα radiation source. An Axis Ultra DLD X-ray photoelectron

spectrophotometer (XPS) equipped with an Al Kα excitation source (1486.69 eV) was used to

record compositional information of all the samples. The energy step size of the XPS was 1 eV

for the survey scans and 0.1 eV for the fine scans. Raman spectra were conducted on a Horiba

Micro Raman HR Evolution System. High-resolution transmission electron microscopy (HR-

TEM) of the samples were characterized using a field-emission transmission electron microscope

(FE-TEM, JEM-2010F, JEOL, Japan), which was operated at an accelerating voltage of 200 kV.

Water contact angle measurement was done using a VCA Optima series. Compression test was

done by Lloyd ez10. The four-point probe measurement method was carried out in the 2638A

Hydra Series III data acquisition unit, with the precision of 0.0001 Ohm. Catalytic behaviour of

as-prepared integrated electrode was recorded by VMP3 electrochemical workstation (Bio-logic

Inc.). All the measurements were conducted based on a three-electrode system consisting of a

self-fabricated electrode as a working electrode, a platinum plate as the counter electrode and a

Hg/HgO electrode (1M KOH) as a reference electrode. Before actual data recording, the working

electrode was scanned 20 cycles for LSV to obtain steady graphs in 1M KOH with a scan speed

of 5 mV/s at room temperature. In order to calculate ECSA, the electrochemical double-layer

capacitance (EDLC) of working electrodes was measured by CV scans at different scan rates (5,

10, 20, 40, 80 mV/s). EDLC was therefore calculated by plotting graphs of scan rate versus

current density at a particular potential against the reference electrode. Tafel slopes of the

experiments were all derived from LSV curves. The as fabricated and surface modified

(immersed in 50% HNO3 aqueous solution) graphene foams were placed in a punched

polystyrene foam (thermal conductivity ≈0.04 Wm/k) with the graphene foams inserted in the

opening. The entire structure floats on the surface of the water with only the bottom side in direct
contact of water. The experiments were typically conducted at an ambient T of ≈25°C and a

humidity of ≈ 41% in a petri dish. The solvent absorption capability of the as-fabricated

graphene foam was tested by absorbing the red- stained Hexane from water. Oil absorption

performance was investigated by absorbing cooking oil and soybean oil from water.

References
[1] K. S. Novoselov, V. Fal, L. Colombo, P. Gellert, M. Schwab, K. Kim, Nature 2012, 490,

192.

[2] K. Liu, Y.-M. Chen, G. M. Policastro, M. L. Becker, Y. Zhu, ACS Nano 2015, 9, 6041.

[3] a) Z. Qin, G. S. Jung, M. J. Kang, M. J. Buehler, Sci. Adv. 2017, 3, e1601536; b) Y. Jia, L.

Zhang, G. Gao, H. Chen, B. Wang, J. Zhou, M. T. Soo, M. Hong, X. Yan, G. Qian, Adv.

Mater. 2017, 29, 1700017; c) N. Xu, X. Hu, W. Xu, X. Li, L. Zhou, S. Zhu, J. Zhu, Adv.

Mater. 2017, 29, 1606762; d) S.-H. Bae, Y. Lee, B. K. Sharma, H.-J. Lee, J.-H. Kim, J.-

H. Ahn, Carbon 2013, 51, 236.

[4] A. Obraztsov, E. Obraztsova, A. Tyurnina, A. Zolotukhin, Carbon 2007, 45, 2017.

[5] a) W. S. V. Lee, E. Peng, D. C. Choy, J. M. Xue, Journal of Materials Chemistry A 2015,

3, 19144; b) X. F. Jiang, R. Li, M. Hu, Z. Hu, D. Golberg, Y. Bando, X. B. Wang, Adv.

Mater. 2019, 31, 1901186.

[6] a) A. N. Obraztsov, Nat. Nanotechnol. 2009, 4, 212; b) M. Segal, Nat. Nanotechnol.

2009, 4, 612; c) N. Yousefi, X. Lu, M. Elimelech, N. Tufenkji, Nat. Nanotechnol. 2019,

14, 107.

[7] a) K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-

Y. Choi, B. H. Hong, Nature 2009, 457, 706; b) A. Reina, X. Jia, J. Ho, D. Nezich,
H. Son, V. Bulovic, M. S. Dresselhaus, J. Kong*, Nano Lett. 2009, 9, 3087; c) H. Cao,

Q. Yu, R. Colby, D. Pandey, C. Park, J. Lian, D. Zemlyanov, I. Childres, V. Drachev,

E. A. Stach, J. Appl. Phys. 2010, 107, 044310.

[8] a) Y. Meng, Y. Zhao, C. Hu, H. Cheng, Y. Hu, Z. Zhang, G. Shi, L. Qu, Adv. Mater.

2013, 25, 2326; b) Y. Wang, L. Wang, T. Yang, X. Li, X. Zang, M. Zhu, K. Wang, D.

Wu, H. Zhu, Adv. Funct. Mater. 2014, 24, 4666; c) K. P. Bera, G. Haider, M. Usman,

P. K. Roy, H. I. Lin, Y. M. Liao, C. R. P. Inbaraj, Y. R. Liou, M. Kataria, K. L. Lu,

Adv. Funct. Mater. 2018, 28, 1804802.

[9] a) O. Balci, N. Kakenov, E. Karademir, S. Balci, S. Cakmakyapan, E. O. Polat, H.

Caglayan, E. Özbay, C. Kocabas, Sci. Adv. 2018, 4, 1749; b) C. T. Phare, Y.-H. D. Lee, J.

Cardenas, M. Lipson, Nat. Photonics 2015, 9, 511.

[10] a) G.-L. Chai, K. Qiu, M. Qiao, M.-M. Titirici, C. Shang, Z. Guo, Energy Environ. Sci.

2017, 10, 1186; b) C. Tang, B. Q. Li, Q. Zhang, L. Zhu, H. F. Wang, J. L. Shi, F. Wei,

Adv. Funct. Mater. 2016, 26, 577.

[11] a) J. Yun, Y. Lim, G. N. Jang, D. Kim, S.-J. Lee, H. Park, S. Y. Hong, G. Lee, G. Zi, J.

S. Ha, Nano Energy 2016, 19, 401; b) S. Goossens, G. Navickaite, C. Monasterio, S.

Gupta, J. J. Piqueras, R. Pérez, G. Burwell, I. Nikitskiy, T. Lasanta, T. Galán, Nat.

Photonics 2017, 11, 366.

[12] a) P. B. Pawar, S. Saxena, D. K. Badhe, R. P. Chaudhary, S. Shukla, Sci. Rep. 2016, 6,

21150; b) S. Ramirez, K. Chan, R. Hernandez, E. Recinos, E. Hernandez, R. Salgado,

A. Khitun, J. Garay, A. Balandin, Mater. Des. 2017, 118, 75; c) X. Hu, W. Xu, L. Zhou,

Y. Tan, Y. Wang, S. Zhu, J. Zhu, Adv. Mater. 2017, 29, 1604031.

[13] Y. Zhang, Y. Huang, T. Zhang, H. Chang, P. Xiao, H. Chen, Z. Huang, Y. Chen, Adv.

Mater. 2015, 27, 2049.


[14] a) Z. Chen, W. Ren, L. Gao, B. Liu, S. Pei, H.-M. Cheng, Nat. Mater. 2011, 10, 424; b)

H. Bi, F. Huang, J. Liang, Y. Tang, X. Lü, X. Xie, M. Jiang, J. Mater. Chem. A 2011, 21,

17366; c) X. Xiao, T. E. Beechem, M. T. Brumbach, T. N. Lambert, D. J. Davis, J. R.

Michael, C. M. Washburn, J. Wang, S. M. Brozik, D. R. Wheeler, ACS Nano 2012, 6,

3573.

[15] a) J. Zhao, H. Lai, Z. Lyu, Y. Jiang, K. Xie, X. Wang, Q. Wu, L. Yang, Z. Jin, Y. Ma,

Adv. Mater. 2015, 27, 3541; b) Z. Lyu, D. Xu, L. Yang, R. Che, R. Feng, J. Zhao, Y. Li,

Q. Wu, X. Wang, Z. Hu, Nano Energy 2015, 12, 657; c) G. Li, J. Sun, W. Hou, S. Jiang,

Y. Huang, J. Geng, Nat. Commun. 2016, 7, 10601; d) H. Wang, K. Sun, F. Tao,

D. J. Stacchiola, Y. H. Hu, Angewandte Chemie International Edition 2013, 52, 9210; e)

H. Bi, I. W. Chen, T. Lin, F. Huang, Adv. Mater. 2015, 27, 5943.

[16] L. Shi, K. Chen, R. Du, A. Bachmatiuk, M. H. Rümmeli, K. Xie, Y. Huang, Y. Zhang,

Z. Liu, J. Am. Chem. Soc. 2016, 138, 6360.

[17] D. M. Knotter, J. Am. Chem. Soc. 2000, 122, 4345.

[18] a) E. García- Tuñon, S. Barg, J. Franco, R. Bell, S. Eslava, E. D'Elia, R. C. Maher, F. Guitian,
E. Saiz, Adv. Mater. 2015, 27, 1688; b) Q. Zhang, F. Zhang, S. P. Medarametla, H. Li, C.
Zhou, D. Lin, Small 2016, 12, 1702.

[19] Z. Pei, H. Li, Y. Huang, Q. Xue, Y. Huang, M. Zhu, Z. Wang, C. Zhi, Energy Environ.

Sci. 2017, 10, 742.

[20] a) F. Kotz, K. Arnold, W. Bauer, D. Schild, N. Keller, K. Sachsenheimer, T. M. Nargang,

C. Richter, D. Helmer, B. E. Rapp, Nature 2017, 544, 337; b) D. T. Nguyen,

C. Meyers, T. D. Yee, N. A. Dudukovic, J. F. Destino, C. Zhu, E. B. Duoss, T. F.

Baumann, T. Suratwala, J. E. Smay, Adv. Mater. 2017, 29, 1701181.

[21] D. Zhang, E. Peng, R. Borayek, J. Ding, Adv. Funct. Mater. 2019, 29, 1807082.
[22] S. Xiao, W. Zhu, P. Liu, F. Liu, W. Dai, D. Zhang, W. Chen, H. Li, Nanoscale 2016, 8,

2899.

[23] M. Mecklenburg, A. Schuchardt, Y. K. Mishra, S. Kaps, R. Adelung, A. Lotnyk, L.

Kienle, K. Schulte, Adv. Mater. 2012, 24, 3486.

[24] Y. Jiang, Z. Xu, T. Huang, Y. Liu, F. Guo, J. Xi, W. Gao, C. Gao, Adv. Funct. Mater.

2018, 28, 1707024.

[25] R. M. Hensleigh, H. Cui, J. S. Oakdale, C. Y. Jianchao, P. G. Campbell, E. B. Duoss,

C. M. Spadaccini, X. Zheng, M. A. Worsley, Mater. Horizons 2018, 5, 1035.

[26] J. Zou, J. Liu, A. S. Karakoti, A. Kumar, D. Joung, Q. Li, S. I. Khondaker, S. Seal, L.

Zhai, ACS Nano 2010, 4, 7293.


Supporting Information

Figure S1. Flow ramp cycling of silica resin (tested for 50 cycles, lasted for 48 hours).
Figure S2. Thermogravimetric Analysis (TGA) graph of silica resin.
Figure S3. SEM images of silica sintered body under different sintering condition: (a) dwelling time: 5
hours, sintering temperature: 1250 ℃; (b) dwelling time: 5 hours, sintering temperature: 1350 ℃; (c)
dwelling time: 5 hours, sintering temperature: 1450 ℃ (Inserted picture on the bottom right showed
sintering cracks under this condition); (d) dwelling time: 10 hours, sintering temperature: 1350 ℃.
(Inserted picture on the bottom left showed uniform sintering under this condition)
Figure S4. SEM images of 3D graphene foam grown under different flow condition: (a) CH4 flow rate: 20
sccm, operation time: 1 hour; (b) CH4 flow rate: 10 sccm, operation time: 1 hours; (c) CH4 flow rate: 20
sccm, operation time: 2 hours.
Figure S5. Fabrication of crack-free and phase-pure hierarchical designed graphene foam with
controllable grain size and shape. (a) Scanning electron microscopy (SEM) of the colloidal silica resin
used in this study; (b -c)Sintering of the spin-coated resin and 3D printed fine structure with same
feature size; (d-e) The 3D graphene assemblies from conventional casting(d) and 3D-printed template
(e) lead to different product features after wet etching; (f) SEM image of the microstructure of the
fully processed 3D graphene foam; (g) Energy-dispersive X-ray spectroscopy (EDX) of the fully
processed 3D graphene foam; (h) Cross-sectional elemental line scan of CVD grown solid silica cube
(0.7 x 0.7 x 0.7 cm3) by casting and molding after etching (Direction: from the left (I) to the right
(III)); (i) Cross-sectional elemental line scan of CVD grown silica mesh thread (400 x 500 x 10000
μm3) by 3D printing from the left to the right.
Figure S6. Optical images of as-fabricated 3D lightweight graphene assemblies with various structures (I),
showing twisting (II), rolling (III), bending (IV) motion.
Figure S7. Strength against density in the material property chart.
Figure S8. Tafel slope of all samples for OER reaction.
Table S1. Representative works on fabricating of 3D graphene monolith
Young's
BET surface Conductivity/ Strength/ Compressive
Sample Methods Modulus/
area/ m2 g-1 S cm-1 kPa strain
kPa

3D-printed gyroid
3D printed silica 2.39
graphene foam 994.2 95 55% 239.7
derived (DLP) (~18 mg/cm3)
(This work)

3D printed
0.154
GO[1] (Robocasting; 84.1 50% 174.6
(10 mg/cm3)
Areogel)

3D printed 130
GO + PMA + 0.4
(Robocasting; 20 50% (16
PEG[2] (6 mg/cm3)
Areogel) mg/cm3)

3D printed 0.64
GO[3] 130
(DLP; Areogel) (92 mg/cm3)

Glucose-strutted
solution-based Foldable
5.29 after
preparation after
Graphene Paper[4] 340 reduction
(Mechanical compressio
(65 mg/cm3)
compression; n but brittle
Aerogel)
Solution-based
Graphene
self-assembly 370
monolith[5]
preparation

Solution-based
0.26 16.6
CNT Areogel [6]
self-assembly 580 33%
(11.6 mg/cm3) (95%)
preparation

Prolysis of
Porous Carbon
polymer (silica 830
Skelton[7]
powder-assisted)

Zinc-assisted
Graphene 1.6
solid-state 2020 133
monolith[8] (32 mg/cm3)
pyrolysis
Table S2. Representative works on carbon-materials for desalination
Materials Surface Power density Efficiency
3D-printed gyroid graphene foam (This work) Hydrophobic 1 84.50
3D-printed treated gyroid graphene foam (This work) Hydrophilic 1 93.00
Carbonized mushrooms[9] Hydrophilic 1 78.00
Vertically aligned graphene sheets membrane[10] 1 86.50
Graphene aerogel[11] Hydrophilic 1 53.60
CNT/macroporous silica[12] Hydrophobic 1 82.00
3D-printed porous concave structure[13] Hydrophilic 1 85.60
3D cross-linked honeycomb graphene foam[14] Hydrophilic 1 87.00
Reference:
[1] Q. Zhang, F. Zhang, S. P. Medarametla, H. Li, C. Zhou, D. Lin, Small 2016, 12, 1702.

[2] E. García-Tuñon, S. Barg, J. Franco, R. Bell, S. Eslava, E. D'Elia, R. C. Maher, F.

Guitian, E. Saiz, Adv. Mater. 2015, 27, 1688.

[3] R. M. Hensleigh, H. Cui, J. S. Oakdale, C. Y. Jianchao, P. G. Campbell, E. B. Duoss, C.

M. Spadaccini, X. Zheng, M. A. Worsley, Mater. Horizons 2018, 5, 1035.

[4] W. S. V. Lee, E. Peng, D. C. Choy, J. M. Xue, J. Mater. Chem. A 2015, 3, 19144.

[5] W. Lv, C. Zhang, Z. Li, Q.-H. Yang, The journal of physical chemistry letters 2015, 6,

658.

[6] J. Zou, J. Liu, A. S. Karakoti, A. Kumar, D. Joung, Q. Li, S. I. Khondaker, S. Seal, L.

Zhai, ACS Nano 2010, 4, 7293.

[7] Z. Pei, H. Li, Y. Huang, Q. Xue, Y. Huang, M. Zhu, Z. Wang, C. Zhi, Energy Environ.

Sci. 2017, 10, 742.

[8] X. F. Jiang, R. Li, M. Hu, Z. Hu, D. Golberg, Y. Bando, X. B. Wang, Adv. Mater. 2019,

31, 1901186.

[9] N. Xu, X. Hu, W. Xu, X. Li, L. Zhou, S. Zhu, J. Zhu, Adv. Mater. 2017, 29, 1606762.

[10] P. Zhang, J. Li, L. Lv, Y. Zhao, L. Qu, ACS Nano 2017, 11, 5087.

[11] Y. Fu, G. Wang, T. Mei, J. Li, J. Wang, X. Wang, ACS Sustainable Chemistry &

Engineering 2017, 5, 4665.

[12] Y. Wang, L. Zhang, P. Wang, ACS Sustainable Chemistry & Engineering 2016, 4, 1223.

[13] Y. Li, T. Gao, Z. Yang, C. Chen, W. Luo, J. Song, E. Hitz, C. Jia, Y. Zhou, B. Liu, Adv.

Mater. 2017, 29, 1700981.

[14] Y. Yang, R. Zhao, T. Zhang, K. Zhao, P. Xiao, Y. Ma, P. M. Ajayan, G. Shi, Y. Chen,

ACS Nano 2018, 12, 829.

S-ar putea să vă placă și