Hierarchically Porous Graphitic Carbon with Simultaneously High

Oct 18, 2017 - This HPC-G material was explored for use both as a supercapacitor electrode and for oil adsorption, two applications that require eithe...
0 downloads 14 Views 3MB Size
Subscriber access provided by Universitaetsbibliothek | Johann Christian Senckenberg

Article

Hierarchically Porous Graphitic Carbon with Simultaneously High Surface Area and Colossal Pore Volume Engineered via Ice Templating Luis Estevez, Venkateshkumar Prabhakaran, Adam L Garcia, Yongsoon Shin, Jinhui Tao, Ashleigh M Schwarz, Jens Darsell, Priyanka Bhattacharya, Vaithiyalingam Shutthanandan, and Ji-Guang Zhang ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.7b05085 • Publication Date (Web): 18 Oct 2017 Downloaded from http://pubs.acs.org on October 19, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Nano is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

43x23mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

Hierarchically Porous Graphitic Carbon with Simultaneously High Surface Area and Colossal Pore Volume Engineered via Ice Templating Luis Estevez*a, Venkateshkumar Prabhakarana, Adam L Garciaa, Yongsoon Shina, Jinhui Taoa, Ashleigh M Schwarza, Jens Darsella, Priyanka Bhattacharyaa,b, Vaithiyalingam Shutthanandana, and Ji-Guang Zhang*a a

Pacific Northwest National Laboratory, 902 Battelle Boulevard, P.O. Box 999, Richland, WA

99352 (USA) and bEnergy Technologies and Materials Division, University of Dayton Research Institute KEYWORDS: hierarchically porous carbon, graphitic, ice templating, supercapacitors, water treatment, pore volume

ABSTRACT: Developing hierarchical porous carbon (HPC) materials with competing textural characteristics such as surface area and pore volume in one material is difficult to accomplish— particularly for an atomically ordered graphitic carbon. Herein we describe a synthesis strategy to engineer tunable HPC materials across micro- meso- and macroporous length scales, allowing the fabrication of a graphitic HPC material (HPC-G) with both very high surface area (>2500

ACS Paragon Plus Environment

1

Page 3 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

m2/g) and pore volume (>11 cm3/g), the combination of which has not been attained previously. The mesopore volume alone for these materials is up to 7.53 cm3/g, the highest ever reported— higher than even any porous carbon’s total pore volume, which for our HPC-G material was >11 cm3/g. This HPC-G material was explored for use as both a supercapacitor electrode and for oil adsorption; two applications that require either high surface area or large pore volume—textural properties that are typically exclusive to one another. We accomplished these high textural characteristics by employing ice templating not only as a route for macroporous formation, but as a synergistic vehicle that enabled the significant loading of the mesoporous hard template. This design scheme for HPC-G materials can be utilized in broad applications, including electrochemical systems such as batteries and supercapacitors, sorbents, as well as catalyst supports—particularly supports where a high degree of thermal stability are required.

Hierarchical porous carbon (HPC) materials have garnered much interest recently.1-3 The specific ability to utilize a multi-modal pore size distribution across multiple length scales has been shown to be useful in various applications. The inherent characteristics of carbon (relatively inert, good electrical conductivity and low density) combines well with a tunable hierarchal porosity for various energy based applications that utilize porous carbons for electrode materials such as supercapacitors,4,5,45,46 batteries,6,7 and fuel cells.8,9 HPCs have also been successfully utilized for environmental applications such as sorbents for CO2 capture,10,11 as materials for water treatment in the form of contaminant sorbents12 and capacitive deionization;13 as well as scaffolds for the impregnation of catalyst materials.14,15

ACS Paragon Plus Environment

2

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

The advantages of multimodal pore sizes in HPC materials differ depending on the application(s) for which they are utilized. For energy storage applications, researchers have found that a hierarchical morphology can be of benefit in lowering the kinetic barriers for electrolyte penetration. This can be easily seen in electric double-layer capacitor (EDLC) based supercapacitors,16,17 where the high surface area sites are typically made up of smaller micropores (defined as 1500 m2/g) are usually associated with smaller pores (3 cm3/g) require larger sized pores— typically at least tens of nanometers. Such HPCs have proven to be beneficial when utilized as catalyst supports in direct methanol fuel cells, where high surface area smaller pores are needed for catalyst loading, but at the same time larger pores are required for efficient fuel transport.20 Many researchers have recently utilized dual high surface area/pore volume HPC materials as solid supports for the impregnation of amines with excellent results for CO2 capture.11

ACS Paragon Plus Environment

3

Page 5 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Additionally, the ability to generate HPCs with macropores (>50 nm) has been utilized successfully in flow battery electrodes21 and water treatment via capacitive desalination with flow-through electrodes.22 However, it is extremely rare to find HPC materials with both exceptionally large surface area (>2500 m2/g) and large pore volume (>5 cm3/g), together in one HPC material. Atomic morphology, such as the extent of the ordering within the HPC material (or level of graphitization) is extremely important as well.23,24 The level of graphitization for porous carbon materials has been well documented to be, in large part, due to the carbon precursor utilized and the final pyrolysis temperature employed.25 Electrical conductivity and improved thermal and chemical stability26 of the porous carbons tend to scale with the level of graphitization,23,24 thus these carbons are typically preferred in electrochemical applications. A good example of this required stability is found in platinum supports utilized for catalysis, where the carbon materials require a more graphitic carbon to offset the Pt induced carbon oxidation at lower temperatures.27,28 Herein, we report a rational approach for synthesizing a graphitic HPC system (labeled as HPC-G) and utilizing it as a tunable materials platform with exquisite morphological control of the HPC that enables the realization of extremely high surface area (>2900 m2/g) and colossally large pore volume (>11 cm3/g) values, never before seen together in one HPC material (graphitic or otherwise). We also show how we utilized ice templating to synergistically achieve these textural characteristics. For comparison of the graphitic structure in the HPC-G materials platform, a hard carbon (more atomically disordered) material was also synthesized with similar textural properties to the HPC-G materials and denoted as HPC-S. The HPC-S series of materials is similar to our previous work29 and details of the synthesis and characterization can be found in

ACS Paragon Plus Environment

4

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

the supporting information, whereas the differences in the synthesis technique for the HPC-G materials can be found in the Methods section.

RESULTS AND DISCUSSION Synthesis and Characterization for the HPC-G Materials. For the synthesis of our HPC-G materials, polyacrylonitrile (PAN) was chosen as the carbon precursor. PAN is in many ways an ideal carbon precursor, as it has a high degree of graphitization,30-32 even at a relatively lower graphitization temperature of ~1600°C.31 This high level of graphitization is not only important in electrochemical applications due to its good electrical conductivity and stability, but it is also important in the carbon fiber industry, where the graphitic ordering results in high strength and thermally stable fibers; which enables PAN to dominate the industry.30,31 PAN is also an ideal precursor for its greater carbon yield and lower expense when compared to other high graphitizing precursors such as pitch. Despite the advantages realized in carbon fiber synthesis; one of the major drawbacks in utilizing PAN for the generation of nanometer sized pores is that it is relatively insoluble.31 This limits PAN to systems involving either dimethyl sulfoxide (DMSO) or dimethylformamide when attempting to utilize it in colloidal silica hard templating strategies. Silica suspension based templating strategies typically utilize poorer graphitizing carbon precursors and often complex or expensive strategies must be employed to keep the silica nanoparticles from aggregating during the synthesis.33 Usually when silica is employed, it is in the form of an already nanostructured material that must also be synthesized, or often requires expensive/difficult self-assembly techniques.32,33 Using electrospinning techniques can mitigate these issues,34 but lower silica loadings with respect to the PAN typically results in reduced porosity and surface areas. In this work, we discovered that this was a crucial advantage of

ACS Paragon Plus Environment

5

Page 7 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

utilizing ice templating alongside a colloidal silica hard template, in that it allows for the heavy loading of silica to the PAN precursor. We discovered that ice templating was not only a macroporous pore former,29,35,36 but it worked synergistically with colloidal silica suspensions to allow for the facile, high loading of well dispersed silica nanoparticles within a PAN matrix— crucial for achieving these graphitic HPC-G materials with such high combined surface area and pore volume values. To realize this, we first replaced the water in the silica suspension with DMSO (details in the Supporting Information). Next we dissolved the PAN into the silica DMSO suspension. Typically, PAN/DMSO solutions are limited to ~10 wt. % PAN, as above that threshold, solutions become quite viscous and difficult to work with. By using ice templating, we were able to heavily dilute the mixture of silica nanoparticles and PAN in DMSO, as needed, to ensure heavy silica to PAN loading and subsequent facile removal of the DMSO via lyophilization. This heavy silica loading, (~67 wt. %) is what allowed us to achieve such impressively high porosity, surface area and pore volume values. This is especially apparent after calculating the mass of the silica in relation to the pyrolyzed carbon, roughly 80 wt. %, when estimating a 50% carbon yield.31 Furthermore, the capacity of ice templating to “lock in” a well dispersed mixture by quick and easy freezing via liquid nitrogen,29,35 allowed us to lock in our well dispersed, heavily loaded, silica within the very stable solid PAN structure after lyophilizing. Though the macroporosity created via ice templating is easily tunable36,37 and we have shown in previous work29 that the mesoporous pore size distribution (PSD) is also adjustable (see Supporting Information, S6 for PSD as a function of hard template loading), we concentrated on utilizing activation to increase the already high surface area and pore volume of our HPC-G materials. We employed a 12 nm colloidal silica template at a silica-to-PAN mass ratio of 2:1,

ACS Paragon Plus Environment

6

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

resulting in the HPC-G-12-21 sample. An alternative HPC-G example with a larger colloidal silica template (~115 nm) and resultant porous structure is presented in the Supporting Information (Figure S7) as a proof of concept towards this facile tunability. The HPC-G-12-21 samples underwent physical activativation38 resulting in extremely high textural properties of both specific surface area (2933 m2/g) and pore volume (11.23 cm3/g) together in one material. Scanning electron microscopy or SEM (Figure 1a), reveals the typical macroporous structure present in the HPC-G materials platform. Transmission electron microscopy (TEM) images (Figure 1b) of the HPC-G-12-21 sample show a mesoporous PSD distribution centered at ~15 nm, consistent with the colloidal silica particle size. More significantly, the high resolution TEM (HRTEM) images confirm the presence of a graphitic structure. Figure 1c shows a typical higher magnification HRTEM image of the individual mesopores for the HPC-G material. The image reveals the mesopore walls are made up of layers of graphitic sheets spaced ~0.34 nm apart (Figure 1d), consistent with a more ordered, graphitic structure. The TEM images for the heavily activated HPC-G-12-21-5h carbon (Figure 1e-f) reveal the mesoporous PSD broadens out to ~20 nm, in good agreement with the nitrogen porosimetry data (vide infra); while the HRTEM image (Figure 1f), also shows a more ordered and graphitic structure when compared to the HPC-S materials (Figure S1c, e)—albeit with thinner mesoporous walls when compared to the unactivated HPC-G-12-21 carbon. This graphitic structure for the HPC-G materials is noteworthy particularly when combined with the textural properties for the activated HPC-G-1221 materials (Table 1). Such high surface areas are rare for porous graphitic carbons; with most cases in the literature being below 2000 m2/g.23-25 This high specific surface area (SSA) combined with the massively large measured pore volume values displayed by our HPC-G materials has not been seen in the literature until now.

ACS Paragon Plus Environment

7

Page 9 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 1. Results of electron microscopy for the HPC-G-12-21 sample reveals (a) the macroporous structure as seen by SEM and the (b) mesopores present in the macroporous walls under TEM, as well as (c,d) the graphitic ordering under HRTEM. The heavily activated HPCG-12-21-5h sample shows (e) similar mesopores present under TEM and (f) graphitic ordering via HRTEM. Table 1. Textural characteristics, level of graphitization and oil sorption data for various porous carbons Porous carbon

BET SSA [m2/g]

Micropore SSA [m2/g]a

Total pore volume [cm3/g]b

Micropore volume [cm3/g]a

Mesopore volume [cm3/g]c

D band FWHM [cm-1]

G band FWHM [cm-1]

Oil sorption [g/gcarbon]

HPC-G-1000C

652

83

2.01

0.042

1.49

-

-

-

HPC-G-12-21

619

74

2.47

0.035

1.70

90

67

45

HPC-G-12-21-3h

1831

259

7.03

0.131

4.85

-

-

-

ACS Paragon Plus Environment

8

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

HPC-G-12-21-5h

2933

488

11.51

0.247

7.53

85

77

65

HPC-S-4-21

1292

187

4.89

0.098

4.40

180

68

25

HPC-S-4-21-10h

2675

465

10.84

0.242

7.91

165

79

-

Norit AC

1948

389

1.14

0.209

0.61

120

69

7

KB-600JD

1477

43

4.75

0.014

2.24

140

78

20

a

Microporous surface area determined via the t-plot method, b total pore volume obtained via nitrogen uptake at the single point P/Po value of ~0.995 and c mesopore volume was calculated by selecting the BJH cumulative pore volume value at a pore size of 50 nm and subtracting the t-plot micropore volume.

Nitrogen porosimetry was also used to better elucidate the morphology present in the HPC-G12-21 samples during various stages of synthesis. Figure 2 shows the nitrogen porosimetry data for the HPC-G-12-21 samples before high temperature pyrolysis (HPC-G-1000C) and after the ensuing high temperature pyrolysis (HPC-G-12-21), as well as after later increasing levels of activation. As can be seen in Figure 2a, the HPC-G-1000C sample’s PSD (by volume) reveals mesopores centered at ~15 nm. After pyrolysis at 1600°C, the PSD broadens slightly to larger mesopores, increasing the mesopore volume. Furthermore, there is a smaller 2-3 nm peak also present in the PSD (Figure 2a-b) in addition to the 15 nm peak. Upon activation of the HPC-G12-21 sample, that smaller 2-3 nm peak becomes more prominent, while another peak centered at ~8 nm emerges from the shoulder of the unactivated HPC-G-12-21 sample. There are also micropores that contribute to the higher SSA values that, while beyond the limit of the BarrettJoyner-Halanda (BJH) analysis, can be measured by way of the calculated t-plot micropore surface area. Figure 2c reveals that roughly 82% (9.21 cm3/g) of the pore volume is from pores ≤100 nm and roughly 70% of the pore volume arises from pores sized ≤50 nm (~7.78 cm3/g). Removing the pore volume calculated via the t-plot analysis indicates that the mesopore volume is 7.53 cm3/g—extremely high, especially considering the graphitic nature of the material.

ACS Paragon Plus Environment

9

Page 11 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 2. Nitrogen porosimetry data for the HPC-G materials at varying levels of activation showing (a) pore size distribution curves by volume and (b) surface area, as well as (c) cumulative pore volume as a function of pore size; all based on BJH theory. The (d) isotherms for the various levels of activation are also shown. Though the HRTEM images (vide supra) reveal good graphitic ordering, bulk characterization techniques such as Raman spectroscopy (Figure 3) were also employed to explore the nature of the chemical bonds for the various porous carbons. As can be seen from the Raman spectra data, all the carbons tested had the characteristic D (at ~1350 cm-1) and G (at ~1582 cm-1) bands that are present in many carbon materials.39,40 The G band (named after graphite) is common to all

ACS Paragon Plus Environment

10

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

sp2 carbon systems, whereas the D band is indicative of a hybridized vibrational mode related to the graphene edges and thus reveals the presence of disorder in the graphene structure. Though the ratio of G band to D band can be a good indicator of the level of ordering/graphitization in carbon materials (as shown in the graphite powder sample), a non-porous microcrystalline natural graphite sample will have a much larger D band intensity when compared to its G band,39 due to the small crystal grains in the structure. Thus, despite being naturally occurring graphite, these small graphitic domains can lead a misdiagnosis of the material being non-graphitic if one only analyzes the ratio of the two bands. Similarly, a highly porous graphitic carbon can be thought to be comprised of “small crystalline domains”, with the well-ordered/graphitic carbon located in between the small nanometer pores. Thus, the narrowness of the G and D peaks (indicating crystallinity) will then provide a good indication of the graphitic ordering for the porous carbons. In addition to the narrow crystalline D and G bands, the presence of a G’ band (~2700 cm-1) will also indicate that the carbon material is graphitic.39,40 Compared to the welldefined and separated D and G bands present in the graphite powder, The HPC-G sample pyrolyzed to 1000°C shows very broad bands with little separation of the two bands, consistent with a carbon having many functional groups, as confirmed via XPS (Supporting Information, Table S3). Table 1 shows the full width at half maximum (FWHM) values (see Supporting Information, Figure S5 for peak fit details) for both G and D bands for the remaining porous carbons. For the G band, the values are all quite similar, ranging from 67-79 cm-1. The D band FWHM values are more telling and indicate a good graphitic ordering for the non-activated and highly activated HPC-G samples (90 cm-1 and 85 cm-1, respectively). Commercial porous carbons were also investigated and found to have a wider D band indicative of a more disordered atomic structure (FWHM values of 120 cm-1 and 140 cm-1 for the Norit AC and KB-600JD

ACS Paragon Plus Environment

11

Page 13 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

samples, respectively). The HPC-S samples, being essentially hard carbons,25,41 have the broadest D bands at 180 cm-1 and 165 cm-1 for the unactivated and heavily activated HPC-S samples, respectively. Also of importance is the G’ band found in the graphite sample (at ~2700 cm-1) which is only present for the two (high temperature pyrolyzed) HPC-G samples, giving further credence to their graphitic/ordered structure.

Figure 3. Micro-Raman spectroscopy data for the various HPC-S and HPC-G materials (both heavily activated and non-activated) as well as commercially available porous carbons Norit AC and KB-600JD. A non-porous graphite power was also included for comparison showing graphite’s telltale D, G and G’ bands at ~1360 cm-1, ~1575 cm-1, and ~2700 cm-1 respectively X-ray diffraction (XRD) analysis (Supporting Information, Figure S3) reveals somewhat broad peaks centered at 2θ = ~24° for all of the samples due to the small crystalline domains in the pores. Of interest, is the highly activated HPC-G material, which has a sharp peak beginning to emerge at 2θ = ~26°, just to the right of the broad peak at 24° which is consistent with the earlier work of Ryoo et al.25 In Ryoo’s work, XRD spectra showed the same small sharp peak, to the

ACS Paragon Plus Environment

12

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

right of the broader peak, as the carbons under investigation began to graphitize. As has been previously discussed, thermal stability for carbon materials is an important property in carbons utilized for catalyst supports, especially when utilizing platinum catalysts in electrochemical convertors. We compared the thermal stability of various samples (Figure S4) including our HPC samples, KB-600JD, Norit AC and a non-porous graphite powder. As expected, the nonporous graphite powder was the most thermally stable, whereas the carbon with the least degree of ordering and extremely large SSA, the HPC-S-4-21-10h sample, was the least thermally stable. The onset of oxidation was determined via thermogravimetric analysis and shows a clear exothermic release upon heating that intensifies with mass loss, as would be expected with oxidation/burning of the carbon for all the samples. The onset value for the graphite powder is 664°C, followed by non-activated HPC-G (646°C), KB (622°C), Norit AC (600°C), highly activated HPC-G (597°C) and finally the highly activated HPC-S sample (490°C). Decreasing oxidation onset values are inversely correlated with increasing surface area for all of the samples, but the HPC-G-12-21-5h sample shows relatively good thermal stability despite its extremely high surface area (>2900 m2/g). The HPC-G-12-21-5h sample, despite having roughly twice the surface area and pore volume of the KB, has an oxidation onset value separated by only ~25°C from the KB sample and within 70°C of the nonporous graphite powder. Such impressive thermal stability at such a high SSA is indicative of a more ordered graphitic structure, further corroborating the same conclusions from the Raman and HRTEM analysis previously discussed. HPC-G Materials with Very High Surface Area and Pore Volume as Supercapacitor Electrodes and Oil Adsorbents. To demonstrate the HPC-G materials’ high specific surface area, the application of a supercapacitor electrode was chosen. Supercapacitors (SCs) store energy via electrical double layer charge accumulation and thus require an electrically

ACS Paragon Plus Environment

13

Page 15 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

conductive material with high specific surface area to achieve higher performance; making nanoporous carbons with high surface area, a natural choice.1,4 Additionally, to demonstrate the HPC-G materials’ colossally high pore volume values and graphitic/hydrophobic nature, we tested the materials as oil adsorbents. Due to the HPC-G’s structure we were required to use two different commercial carbons as baselines for comparison: one with high SSA (Norit AC) and another with high pore volume (KB-600JD). For the supercapacitor electrode application, a solid membrane incorporated with an ionic liquid (IL) based electrolyte, 1-Ethyl-3-methylimidazolium tetrafluoroborate (EMIMBF4) was chosen, in a symmetrical two electrode setup (each at ~3.5 mg/cm2 loading of active material, see Supporting Information for further set-up details). An IL based electrolyte was primarily chosen to measure and compare the specific capacitance at a low rate where the porous carbons’ SSA would dictate most of the SC performance and secondly, at higher rates to see the effects of bulky IL ions in an open hierarchical structure versus a more ramified microporous structure such as for traditional activated carbons (Norit AC). It should be noted however, that such an exploration was limited in scope for this manuscript and the SC electrode was primarily utilized to comparatively demonstrate the very high SSA of our HPC materials alongside another high SSA porous carbon. This is particularly relevant considering the use of solid film membranes being utilized as an electrolyte, compared to the more commonly utilized liquid IL electrolytes employed in the literature.

ACS Paragon Plus Environment

14

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

Figure 4. Comparison of high surface area commercial carbon (Norit AC) and HPC-G-12-21-5h as electrochemical capacitor electrodes. Specific capacitance values were obtained from (a) galvanostatic charge-discharge curves and (b) were plotted as a function of rate (based on active material). (c) GCD curves for both materials at a rate of 0.6 A/g reveals the superior rate performance imbued by the morphology of the HPC-G; further corroborated by (d) cyclic voltammetry curves at a rate of 10 mV/s. Figure 4 shows the results of the supercapacitor experiments. At a very low rate of 0.1 A/g, the galvanostatic charge-discharge (GCD) curves (Figure 4a) reveal stable supercapacitor performance. At such low rates, the bulky IL ions can more easily find the high surface area sites on both the Norit AC and HPC-G electrodes and the SSA plays a greater role in the measured

ACS Paragon Plus Environment

15

Page 17 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

specific capacitance values of the carbons. At 0.1 A/g, the specific capacitance values for the HPC-G-12-21-5h (henceforth referred to as HPC-G for brevity) and Norit AC samples are roughly 90 F/g and 75 F/g respectively—in good agreement with the SSA values for the respective carbons (Table 1). Interestingly, even at the low GCD rate of 0.1 A/g, the Norit AC has a noticeable increased IR drop when compared to that of the HPC-G material (0.23 V and 0.07 V, respectively). We attribute this increased IR drop for the Norit AC material to the increased electrical conductivity from the ordered graphitic nature of the HPC-G material, since differences in morphology are minimized at such a low rate. Figure 4b shows the specific capacitance of the two carbons as a function of rate. When the rate is doubled (0.2 A/g), the HPC-G electrode maintains most of its capacitance at a value of ~88 F/g; which, despite the limitations imposed by the solid electrolyte, performs surprisingly well when compared to other similar IL electrolyte SC systems (Table S4), even with ILs in liquid form (and sometimes with organic solvents added).42 Relative to the HPC-G material, the Norit AC electrode starts to lose capacitance relatively quickly at higher rates culminating in a dramatic drop off in capacitance (~39 F/g or ~52% of the original value) at 0.6 A/g. The HPC-G material retains more of its capacitance (~81 F/g or ~90%) at the same 0.6 A/g rate and beyond, as at 0.8 A/g, we were unable to calculate the capacitance for the Norit AC. At 0.6 A/g, the GCD curves (Figure 4c) for both the Norit AC and HPC-G electrodes show an increased IR drop of 1.32 V and 0.37 V respectively. Thus it is clear that at 0.6 A/g, the ions are difficult to move within the high surface area sorption sites of the Norit AC electrode, whereas they are easy to move within the HPC-G electrode. Additionally, the HPC-G electrodes still maintain good symmetry for its GCD curve at 0.6 A/g, indicative of its better performance at higher rates. This is further corroborated via cyclic voltammetry (Figure 4d) where the more distorted rectangular shape of the Norit is

ACS Paragon Plus Environment

16

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

indicative of the poor ion transport in the purely microporous material. These promising results show the superior benefits of an HPC-G morphology as potential electrodes in IL based SCs. Further testing for long-term cyclic stability performance shows good capacity retention (~92%) at 10,000 cycles, typical for a supercapacitor (Figure S9). Furthermore, though beyond the scope of this work, the ability to control the ultimate carbonization temperature for the HPC-G precursor can be utilized in future work, as at the lower pyrolysis temperature of 1000°C leads to a porous carbon material with abundant oxygen (~3.8 at. %) and nitrogen groups (~6.6 at. %) as shown on Table S3. An intrinsic electrochemical activity of the electrode that contains an HPC material (HPC-G-1000C) evaluate using via cyclic voltammetry showed the presence of additional Faradaic activity and attributed to the presence of oxygen and nitrogen functional groups (See Figure S10). To demonstrate the colossally large void space for the HPC-G materials, they were further investigated as oil sorbents. Recently, many efforts have been made to utilize porous carbon materials for oil adsorption.43,44 Two of the main prerequisites for good oil sorption performance in porous carbons are hydrophobicity and available void space (pore volume) and thus, oil adsorption was chosen as an application to demonstrate the extremely high pore volume of the HPC-G materials. KB-600JD was chosen as the commercial carbon for comparison, due its high pore volume (~4.75 cm3/g) and good hydrophobicity, the latter corroborated by its low oxygen content (1.4 at% via XPS, Table S3). The samples were examined for oil adsorption by utilizing 100 – 200 mg of the porous carbon material in a petri dish, whereupon motor oil was pipetted, drop-wise, onto the carbon until saturated (see Supporting Information for details and optical images in Figure S8). It was found that the KB-600JD sample could adsorb ~20 times its own mass in oil, relatively high when compared to a low pore volume porous carbon such as Norit

ACS Paragon Plus Environment

17

Page 19 of 29

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

AC which could only adsorb ~6 times its weight. The unactivated HPC-S-4-21 sample (measured total BJH pore volume of 4.78 cm3/g) was found to adsorb ~25 times its own weight, slightly higher than the KB-600. For the HPC-G system of materials the oil adsorption results were even higher. The unactivated HPC-12-21 had a very high oil sorption capability of ~45 times its own weight. Note that this high oil adsorption performance comes from a material with a relatively lower pore volume of 2.44 cm3/g. We believe this is due to the relatively high hydrophobicity of the unactivated HPC-G-12-21 material as evidenced by its very low oxygen content (0.5 at% via XPS). After extensive activation, the HPC-G-12-21-5h material is still quite hydrophobic, but the process does impart additional oxygen groups which increases its elemental concentration of oxygen to 2.0 at% (Supporting Information, Table S3). Despite the slight increase in hydrophilicity, the colossally high pore volume of >10 cm3/g combined with a still relatively high hydrophobic nature allows the HPC-G-12-21-5h sample to adsorb 65 times its own weight in oil—quite high considering recent results for similar materials.43,44

CONCLUSIONS In summary, a synthesis strategy for a series of tunable hierarchically porous carbons was developed that exhibit a combination of very high surface area (>2500 m2/g via BET) and large pore volume (>10 cm3/g via BJH) in one material—a combination which has not been previously accomplished. The incorporation of ice templating was found to play a highly significant role in enhancing these advanced textural characteristics. This synthesis allows for the tunability of not only the porous morphology, but also for the adjustability down to the atomic structure, permitting the same high textural characteristics in a graphitic/ordered highly porous carbon material. Utilizing one single HPC-G material, we were able to outperform two separate

ACS Paragon Plus Environment

18

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

commercial carbons, specifically selected for either high specific surface area or pore volume in applications that favor each specific textural property. The potential applications presented in this work demonstrated the and advantageous textural properties of the HPC-G material while the synthesis strategy presented herein leads to an excellent direction in the rational design of highly porous graphitic carbon materials.

METHODS Synthesis of the HPC-S materials follows an optimized route based, in part, on our previous work29 which is addressed in detail in the Supporting Information (SI). Synthesis of the HPC-G materials involved using a triple templating strategy whereby ice templating, a colloidal silica hard template and CO2 activation were employed for the formation and control of macropores (>50 nm), mesopores (2-50 nm) and micropores (