Next Article in Journal
Concrete Slab-Type Elements Strengthened with Cast-in-Place Carbon Textile Reinforced Concrete System
Next Article in Special Issue
A Novel Method for the Detection of SARS-CoV-2 Based on Graphene-Impedimetric Immunosensor
Previous Article in Journal
Investigating the Effectiveness of Nano-Montmorillonite on Asphalt Binder from Rheological, Thermodynamics, and Chemical Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Low-Temperature Heat Output Photoactive Material-Based High-Performance Thermal Energy Storage Closed System

1
College of Materials Science and Engineering, Taiyuan University of Technology, Yingze West Street, Taiyuan 030024, China
2
Institute of Carbon Materials Science, Shanxi Datong University, Xingyun Street, Datong 037009, China
*
Authors to whom correspondence should be addressed.
Materials 2021, 14(6), 1434; https://doi.org/10.3390/ma14061434
Submission received: 27 December 2020 / Revised: 26 January 2021 / Accepted: 25 February 2021 / Published: 16 March 2021

Abstract

:
Designing and synthesizing photothermal conversion materials with better storage capacity, long-term stability as well as low temperature energy output capability is still a huge challenge in the area of photothermal storage. In this work, we report a brand new photothermal conversion material obtained by attaching trifluoromethylated azobenzene (AzoF) to reduced graphene oxide (rGO). AzoF-rGO exhibits outstanding heat storage density and power density up to 386.1 kJ·kg−1 and 890.6 W·kg−1, respectively, with a long half-life (87.7 h) because of the H-bonds based on high attachment density. AzoF-rGO also exhibits excellent cycling stability and is equipped with low-temperature energy output capability, which achieves the reversible cycle of photothermal conversion within a closed system. This novel AzoF-rGO complex, which on the one hand exhibits remarkable energy storage performance as well as excellent storage life span, and on the other hand is equipped with the ability to release heat at low temperatures, shows broad prospects in the practical application of actual photothermal storage.

1. Introduction

With the fast development of society, people’s demand for energy is increasing and the energy issue has now become one of the major problems that human beings need to deal with [1]. Solar energy has the advantages of sufficient reserves, no pollution and economical availability. Efficiently converting and storing solar energy has become an important way to overcome the current energy shortage crisis [2,3,4,5]. Recently, photothermal conversion materials have attracted extensive attention as a new method for storing solar energy [6]. Photothermal conversion materials can store solar energy in chemical bonds through photo-isomerization of units and then releasing the stored energy as thermal energy when exposed to different external stimulus, achieving photothermal conversion within a closed system. Such materials are able to effectively convert light energy into its own chemical bonds and release its stored energy while avoiding the emission of additional greenhouse gases, with the potential to achieve low-cost and large-scale industrial solar storage [7]. However, photothermal conversion materials still have the shortcomings of short storage time, low energy density and inability to achieve energy release under low temperatures, which are key factors limiting its practical application in solar thermal energy storage [8,9].
Owing to its special photoisomerization ability, good structural stability and controllable configuration recoverability, azobenzene and its derivatives with numerous applications [10,11] has received extensive research interest as a kind of photothermal conversion material [12,13]. However, due to the disadvantages of poor storage performance and storage half-life (τ1/2) arising from low isomerization enthalpy (ΔH), azobenzene did not exert its full potential in terms of photothermal conversion and storage [14]. To override the above hurdles, great efforts have been made on the basis of molecular design by introducing different substituents and increasing the interaction between molecules [15,16,17]. Grossman et al. [18] reported azobenzene derivatives with bulky aromatic groups as photoactive chemical heat storage materials. Owing to the introduction of bulky phenyl groups, the solid-state azobenzene derivatives not only improve the energy density but also improve the corresponding thermal stability. Bléger et al. [19] reported o-Fluoroazobenzenes and derivatives which exhibit an unprecedented long half-life owing to the ortho-fluorine substituent which reduces electron density around the –N=N– double bond. Despite great efforts having been made, it is still an intractable problem to apply azobenzene photothermal conversion material to practical energy storage.
Different from freely dispersed azobenzene, many azobenzene carbon materials were formed by introducing azobenzene into high-strength carbon nanomaterials forms many azobenzene carbon nanocomposites [7,20,21] accompanied by a more closely ordered structure, which have excellent storage capacity and life cycle. The templated, structure modified azobenzene enhance the intermolecular interactions while obtaining a more stable and tightly ordered structure, which jointly improved the storage capacity of azobenzene carbon materials [22,23]. In addition, because of the unique 2D structure and broad surface of graphene with numerous applications [24,25] which contributes to high attachment density, the templated azobenzene/graphene nanomaterials show broad prospects in photothermal storage [26]. Unfortunately, azobenzene carbon nanomaterials still have problems such as difficulty in releasing storage heat at low temperatures and the inability to balance energy density and half-life, which limits their further practical application [27,28]. Therefore, how to simultaneously achieve the improvement of storage capacity and life cycle with low-temperature energy output capability is still a key issue in current research.
In this work, we report a novel photothermal conversion material by attaching trifluoromethylated azobenzene (AzoF) to reduced graphene oxide (rGO). The storage capacity and storage life span as well as the cycling stability performance of AzoF-rGO has made great progress. AzoF-rGO exhibits great development potential in recyclable and long term photothermal storage.

2. Materials and Methods

2.1. Materials

3-amino-5-(trifluoromethyl)benzoic acid (99%), 3,5-dimethoxyaniline (99%), sodium nitrite (97%), Na2CO3 (97%) and NaBH4 (97%) were purchased from Aladdin Reagent (Shanghai, China).

2.2. Detailed Synthesis Steps

  • 3-amino-5-(trifluoromethyl)benzoic acid (1.025 g) was dissolved in the HCl solution (50 mL, 0.5 mol·L−1), then NaNO2 (0.380 g) was added and reacted at ice bath for 80 min. After dissolving 3,5-dimethoxyaniline (0.765 g) in water, we slowly added the above mixture to it, adjusted the pH to 7 and reacted it in an ice bath for 4 h. AzoF was obtained after further purification (1.255 g, 68%).
  • GO was synthesized according to the literature reports [29]. First, we used NaOH (1 mol·L−1) solution to change the pH of the GO aqueous solution (300 mL, 0.5 mg·mL−1) to 10, then we reacted it at 90 °C for 4 h with NaBH4 (180 mg) under N2 atmosphere. When the reaction was complete, rGO was obtained by washing the mixture with water multiple times.
  • AzoF (0.738 g) was dissolved in the HCl solution (60 mL, 0.5·mol L−1), then NaNO2 (0.141 g) was slowly added and reacted in an ice bath for 80 min, and the above mixture was slowly added to the rGO solution (62 mL, 1 mg·mL−1). The mixture was first reacted at 0 °C for 4 h and then at 30 °C for 16 h. AzoF-rGO was obtained by purifying the mixture with water and DMF multiple times.

2.3. Characterizations

The FT-IR was gathered from Vertex 70 (Bruker, Karlsruhe, Germany). The XRD was gathered from X‘Pert Pro MPD (PANalytical, Almelo, Holland). Raman spectrum was gathered from LabRAM Aramis (HORIBA, Paris, France). The XPS was gathered from ESCALAB 250Xi (ThermoFisher, Waltham, MA, USA) using C1 s = 284.8 eV for energy calibration procedures, Operation Mode:CAE:Pass Energy 100.0 Ev, software:Thermo Avantage 5.976 and hemispherical energy analyzer were used for the test, the test vacuum was 5 × 10−9 Torr, the sample was fixed on the sample stage with conductive glue, the background was buckled through the smart method, and the energy calibration was performed with gold, silver and copper. The TGA was performed on STA449F5 (NETZSCH, Bavaria, Germany). TEM was gathered from Tecnai F20 (FEI, Hillsboro, Oregon, USA). SEM were gathered from SU8010 (Hitachi, Tokyo, Japan). The UV–Vis absorption spectra was performed on SPECORD 50 PLUS (ANALYTIK JENA, Jena, Germany) in the range of 250~550 nm with the resolution of 0.1 nm. The trans cis transition was introduced by a multiband LED lamp at 365 nm. The cis trans transition was introduced by a multiband LED lamp at 540 nm. The light intensity was gathered from an optical power meter (PL-MW2000, Bofeilai Technology, Beijing, China). The heat storage density was determined through differential scanning calorimetry (DSC, 214 Polyma, NETZSCH, Bavaria, Germany) under N2.

3. Results and Discussion

3.1. Chemical Structure

As shown in Figure 1a, the low-resolution TEM image of rGO exhibited a smooth structure and its electron diffraction exhibited a hexagonal lattice according to Fast Fourier Transform (FFT) patterns within Figure 1b, demonstrating its good crystallinity. Figure 1c shows that the surface of the material became rough, and the electron diffraction spot of AzoF-rGO (Figure 1d) has become a closed loop attributed to the adhesion of AzoF on rGO [30,31]. Furthermore, the SEM of AzoF-rGO (Figure 1f) shows a stacking phenomenon compared with rGO (Figure 1e). This phenomenon not only reduced the distance between adjacent graphene layers but also enhanced the intermolecular interaction, resulting in a growth in the storage capacity as well as τ1/2 of AzoF-rGO [21]. In addition, it can also be concluded that the distance between layers was reduced based on the XRD results (Figure S2). After the reduction of GO, the (0 0 1) diffraction peak at 11.3° disappeared [32] and was replaced by the (0 0 2) diffraction peak at 22.9° of rGO, and the corresponding grain size was 25.51 nm based on Scherrer formula [33]. After attaching AzoF onto rGO, the 2θ of AzoF-rGO has become to 25.2° with the grain size of 22.63 nm, which is consistent with the SEM observation (Figure 1f) [34].
The AzoF-rGO had new peaks of –N=N– (1430 cm−1) and –CF3 (1140 cm−1) compared to rGO [35] according to Figure 2a. Moreover, the FT-IR spectra of AzoF-rGO and AzoF also showed peaks derived from -OH (3298 cm−1) and –C=O (1640 cm−1). It can also be seen from Figure 2a that the wavenumbers of -OH and –C=O of AzoF-rGO show a significant red shift compared to that of AzoF (3204 cm−1 and 1700 cm−1), confirming the formation H-bond of AzoF on rGO [36]. XPS results also proven the successful grafting of AzoF on rGO. In addition, the characteristic peaks of AzoF at 287.5 eV and 292.5 eV corresponding to C–N and C–F bond also appeared in AzoF-rGO (Figure S3) [35]. Additionally, the fact that there were characteristic peaks of –N=N– (400.3 eV) and –CF3 (688.3 eV) in AzoF-rGO also confirmed the successful bonding between AzoF and rGO [35].
The high-density adhesion of AzoF onto rGO nanosheets is inextricably linked to the improvement of the performance of AzoF-rGO. The decomposition of rGO during the whole heating process was linear according to Figure 2d, and its weight loss mainly attributed to the disappearance of oxygen-containing groups [37]. The AzoF was stable before 185 °C, and its weight loss was attributed to self-decomposition. Additionally, the weight loss of AzoF-rGO was caused by the weight loss of AzoF and rGO [27]. Therefore, the attachment density (Ad) of AzoF on rGO after different time reactions can be obtained based on Equation (1) [38].
D g = R p R R p R a × 100 %
where Ra is the residual weight percentage of AzoF, R is the residual weight percentage of AzoF-rGO, Rp is the residual weight percentage of rGO.
Table 1 shows that the attachment density (Ad) was 1/40 after the first reaction and increased to 1/16 after the third reaction. The attachment density can also be obtained based on XPS [39]. It can also be seen from Table 1 that the results obtained by XPS and TGA were almost identical. From the above results, it can be concluded that almost every 16 carbon atoms of rGO correspond to one AzoF after the third reaction, which is better than previous research [21,40]. High adhesion density on the one hand helps to form intermolecular hydrogen bonds, while on the other hand it also enhances intermolecular interactions, which improves the storage performance of AzoF-rGO [41]. In addition, Raman spectroscopy also proved this result. It can also be seen from Figure S4 that the ID/IG value of AzoF-rGO-1 (1.14) and AzoF-rGO-3 (1.18) was much larger than rGO (1.08), which indicates that the crystal structure of rGO has changed after attachment [31], proving the remarkable attachment density of AzoF on rGO.

3.2. Cycling Stability and Storage Performance

The optical properties performance of AzoF and AzoF-rGO was investigated through time-evolved absorption spectra. It can be seen from Figure 3 that AzoF-rGO went through a trans cis isomerization process under 365 nm ultraviolet light irradiation. Compared with AzoF (τ1/2: 195.2 min), AzoF-rGO (τ1/2: 87.7 h) takes more time to complete the isomerization process from cis-isomer to trans-isomer, indicating that AzoF-rGO has better thermal stability than pristine AzoF. The same conclusion can be drawn from the fact that the first-order reversion rate constant (Krev) of AzoF-rGO (3.29 × 10−6·s−1) was much smaller than that of AzoF (1.20 × 10−4·s−1) under dark conditions derived from Equation (2) [21].
ln ( A t A A 0 A ) = k r e v t
where A0 is the absorption intensity of AzoF-rGO and AzoF at metastable state (cis-rich) irradiated by UV light, At is the absorbance of AzoF-rGO and AzoF reversing for “t” time and A is the absorption intensity of AzoF-rGO and AzoF after complete cis-to-trans reversion. Moreover, compared to pristine AzoF (Figure S5), AzoF-rGO exhibited a lower isomerization degree owing to the intermolecular H-bonds and steric hindrance owing to high attachment density, resulting in a better storage performance of this material. Furthermore, the ΔEa value of the cis-isomer of AzoF-rGO (1.05 eV) was higher than that of AzoF (0.94 eV) according to Equation (3) [42], which again proves the formation of intermolecular hydrogen bonds [43].
E a = RTln h l n 2 τ 1 / 2 k B T
where Ea is the activation barrier for cis-to-trans isomerization process, T represents the temperature and τ1/2 represents the half-life. kB, R and h are the Boltzman, universal gas and Plank constants. Additionally, the optical band gap of AzoF-rGO complex was estimated to be ~1.8 eV based on the Tauc formula (Figure S6) [44]. The increase in the stability of the cis-isomer means extension of the life cycle of AzoF-rGO, which is directly related to the large-scale promotion of photoactive chemical heat storage materials.
Similar to the length of the life cycle, whether the controllable heat release under external stimuli can be achieved is critical to the future application value of AzoF-rGO. Figure 3c showed that compared with dark conditions, the irradiation of green light (540 nm) significantly accelerated the recovery process of AzoF-rGO from cis -isomer to trans-isomer. Compared with dark conditions, the result that Krev (7.58 × 10−4·s−1) was significantly larger under green light irradiation also confirmed the conclusion of faster reversion. The same effect can also be achieved by absorbing heat from the external environment according to DSC. The reason for this phenomenon is that the cis-isomer can absorb energy from external stimuli to overcome the energy barrier of configuration reversion isomerization, thereby achieving the purpose of accelerating energy output [45,46]. The above results show that AzoF-rGO has successfully possessed the controllable heat output capability.
The stability of repeated cis ↔ trans configuration transformations of AzoF-rGO and AzoF has also been studied. It can be seen from Figure 4 that both AzoF-rGO and AzoF have no significant decrease in the absorption intensity at 407 nm after repeated irradiation of ultraviolet light (365 nm) and visible light (540 nm) for 50 times, which shows that they have outstanding isomerization stability. The AzoF-rGO can not only be stored for a long time under the premise of ensuring the storage effect, but also can control the output of the stored energy, which is essential for actual photothermal conversion.
The photothermal storage capacity of AzoF and AzoF-rGO was investigated through DSC [7]. All objects were stable between 10–140 °C based on TGA. AzoF and AzoF-rGO released significant heat under the first round of heating stimulation, but no heat was released during the second round according to Figure 5. The above results prove that the research subjects have released all the energy stored through the configuration transformation in the form of heat. Furthermore, most photothermal storage materials start to release the stored energy after 100 °C, while this kind of heat storage material can start energy output at 35 °C, which is a milestone in achieving fast energy output at lower temperatures [7].
It can be seen from Figure 5 that the heat storage density of AzoF-rGO-3 has reached to 386.1 kJ kg−1, which shows a significant increase over AzoF (121.4 kJ kg−1). This is because of the close-packed orderly distribute of AzoF on rGO as a result of high attachment density, which strengthens the intermolecular interaction [23]. In addition, high attachment density also enhances the steric hindrance and promotes the formation of H-bonds, which further increases the photothermal storage capacity [47]. The reason for AzoF-rGO-1 showing less effectiveness compared to the AzoF is the low attachment density, which leads to weak intermolecular interaction and therefore relatively low energy density. Moreover, the heat storage density of AzoF-rGO-3 was also higher than AzoF-rGO-1 and AzoF-rGO-2, which shows that the attachment density was positively correlated with great storage performance.
Similar to heat storage density, power density is also a key element to measure the possibility of practical application of AzoF-rGO. It can be seen from Figure 6 that the power density of AzoF-rGO-3 was 890.6 W kg−1, which shows a huge improvement compared to AzoF (448.6 W kg−1). Furthermore, the power density of AzoF-rGO-3 was also higher than AzoF-rGO-1 and AzoF-rGO-2, which shows that the attachment density is directly related to the heat output performance. It is worth noting that high power density means fast output of energy, which further increases the feasibility of practical application of AzoF-rGO. As shown in Table 2, the performance of AzoF-rGO in many aspects has been greatly improved compared to other similar materials [7,15,48,49]. The above results demonstrate that AzoF-rGO, which not only exhibits remarkable photothermal capacity but also equipped with low temperature energy output capability, has shown great development potential in achieving the goal of efficient photothermal storage.

4. Conclusions

In summary, AzoF-rGO with good photothermal storage performance, outstanding storage lifespan and low-temperature energy output capability has been proven to be a great photothermal conversion material. The formation of hydrogen bonds and the enhancement of intermolecular interactions owing to the high attachment density has simultaneously achieved the improvement of the heat storage density (max. 386.1 kJ kg−1), power density (max. 890.6 W kg−1) and half-life (up to 87.7 h) of AzoF-rGO for photothermal storage. AzoF-rGO also exhibits exceptional cycling stability, which realizes long-term recyclability and efficient and pollution-free utilization of solar energy in a closed system. Furthermore, AzoF-rGO can start energy output at 35 °C, which shows that the goal of low-temperature energy output has been achieved. The above results indicate that AzoF-rGO, combining outstanding photothermal capacity with a long-life cycle as well as low-temperature energy output capability, is a prominent photothermal conversion material with great practical application value.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/14/6/1434/s1, Figure S1: 1H NMR, 13C NMR and 19F NMR spectra of AzoF, Figure S2: XRD patterns of GO, rGO, AzoF-rGO, Figure S3: C1s region in XPS spectra of (a) rGO, (b) AzoF-rGO and (c) N1s region in XPS spectra, Figure S4: Raman spectra, Figure S5: Time-evolved absorption spectra of AzoF and Figure S6: UV-Vis absorption spectra of AzoF-rGO powder at room temperature (25 °C).

Author Contributions

Conceptualization, X.Y. and S.L.; methodology, X.Y.; software, X.Y.; validation, X.W. and Y.W.; formal analysis, X.Y.; investigation, X.Y.; resources, J.Z. (Jin Zhang); data curation, X.Y. and S.L.; writing—original draft preparation, X.Y.; writing—review and editing, S.L.; visualization, J.Z. (Jianguo Zhao); supervision, J.Z. (Jin Zhang); project administration, J.Z. (Jianguo Zhao); funding acquisition, J.Z. (Jin Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 52071192), Shanxi “1331 project” foundation for the construction of collaborative innovation center of graphene industrial application, Key Research and Development Project of Shanxi Province (Grant No. 201803D121122), Research and development projects in key areas of Guangdong Province (Grant No. 2020B0202010004), The Program for Scientific and Technological Innovation of Higher Education Institutions in Shanxi (Grant No. 2020L0477).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in [insert article or Supplementary Material].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armaroli, N.; Balzani, V. The future of energy supply: Challenges and opportunities. Angew. Chem. Int. Ed. Engl. 2007, 46, 52–66. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, X.; Peng, T.; Song, S. Recent advances in dye-sensitized semiconductor systems for photocatalytic hydrogen production. J. Mater. Chem. A 2016, 4, 2365–2402. [Google Scholar] [CrossRef]
  3. Aguirre, M.E.; Zhou, R.; Eugene, A.J.; Guzman, M.I.; Grela, M.A. Cu2O/TiO2 heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O from photocorrosion. Appl. Catal. B Environ. 2017, 217, 485. [Google Scholar] [CrossRef]
  4. Zhou, R.; Guzman, M.I. CO2 Reduction under Periodic Illumination of ZnS. J. Phys. Chem. C 2015, 118, 11649–11656. [Google Scholar] [CrossRef]
  5. Moth-Poulsen, K.; Ćoso, D.; Börjesson, K.; Vinokurov, N.; Meier, S.K.; Majumdar, A.; Vollhardt, K.P.C.; Segalman, R.A. Molecular solar thermal (MOST) energy storage and release system. Energy Environ. Sci. 2012, 5, 8534–8537. [Google Scholar] [CrossRef] [Green Version]
  6. Yoshida, Z.-I. New molecular energy storage systems. J. Photochem. 1985, 29, 27–40. [Google Scholar] [CrossRef]
  7. Kucharski, T.J.; Ferralis, N.; Kolpak, A.M.; Zheng, J.O.; Nocera, D.G.; Grossman, J.C. Templated assembly of photoswitches significantly increases the energy-storage capacity of solar thermal fuels. Nature Chem. 2014, 6, 441–447. [Google Scholar] [CrossRef] [PubMed]
  8. Kanai, Y.; Srinivasan, V.; Meier, S.K.; Vollhardt, K.P.; Grossman, J.C. Mechanism of thermal reversal of the (fulvalene)tetracarbonyldiruthenium photoisomerization: Toward molecular solar-thermal energy storage. Angew. Chem. Int. Ed. Engl. 2010, 49, 8926–8929. [Google Scholar] [CrossRef] [Green Version]
  9. Berthiller, F.; Dall’asta, C.; Corradini, R.; Marchelli, R.; Sulyok, M.; Krska, R.; Adam, G.; Schuhmacher, R. Occurrence of deoxynivalenol and its 3-beta-D-glucoside in wheat and maize. Food Addit. Contam. Part A Chem. Anal. Control. Expo. Risk Assess 2009, 26, 507–511. [Google Scholar] [CrossRef] [Green Version]
  10. Lee, K.M.; Smith, M.L.; Koerner, H.; Tabiryan, N.; Vaia, R.A.; Bunning, T.J.; White, T.J. Photodriven, Flexural-Torsional Oscillation of Glassy Azobenzene Liquid Crystal Polymer Networks. Adv. Funct. Mater. 2011, 21, 2913–2918. [Google Scholar] [CrossRef]
  11. Liu, J.; Bu, W.; Pan, L.; Shi, J. NIR-triggered anticancer drug delivery by upconverting nanoparticles with integrated azobenzene-modified mesoporous silica. Angew. Chem. Int. Ed. Engl. 2013, 52, 4375–4379. [Google Scholar] [CrossRef] [PubMed]
  12. Sun, L.; Gao, F.; Shen, D.; Liu, Z.; Yao, Y.; Lin, S. Rationally designed hyperbranched azopolymer with temperature, photo and pH responsive behavior. Polym. Chem. 2018, 9, 2977–2983. [Google Scholar] [CrossRef]
  13. Wang, G.; Yuan, D.; Yuan, T.; Dong, J.; Feng, N.; Han, G. A visible light responsive azobenzene-functionalized polymer: Synthesis, self-assembly, and photoresponsive properties. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 2768–2775. [Google Scholar] [CrossRef]
  14. Iii, J.O.; Lawrence, J.; Yee, G.G. Photochemical storage potential of azobenzenes. Sol. Energy 1983, 30, 271–274. [Google Scholar]
  15. Han, G.G.D.; Li, H.; Grossman, J.C. Optically-controlled long-term storage and release of thermal energy in phase-change materials. Nat. Commun. 2017, 8, 1446. [Google Scholar] [CrossRef] [Green Version]
  16. Joshi, D.K.; Mitchell, M.J.; Bruce, D.; Lough, A.J.; Yan, H. Synthesis of cyclic azobenzene analogues. Tetrahedron 2012, 68, 8670–8676. [Google Scholar] [CrossRef]
  17. Weston, C.E.; Richardson, R.D.; Haycock, P.R.; White, A.J.; Fuchter, M.J. Arylazopyrazoles: Azoheteroarene photoswitches offering quantitative isomerization and long thermal half-lives. J. Am. Chem. Soc. 2014, 136, 11878–11881. [Google Scholar] [CrossRef]
  18. Cho, E.N.; Zhitomirsky, D.; Han, G.G.; Liu, Y.; Grossman, J.C. Molecularly Engineered Azobenzene Derivatives for High Energy Density Solid-State Solar Thermal Fuels. ACS Appl. Mater. Interfaces 2017, 9, 8679–8687. [Google Scholar] [CrossRef]
  19. Bleger, D.; Schwarz, J.; Brouwer, A.M.; Hecht, S. o-Fluoroazobenzenes as readily synthesized photoswitches offering nearly quantitative two-way isomerization with visible light. J. Am. Chem. Soc. 2012, 134, 20597–20600. [Google Scholar] [CrossRef]
  20. Lennartson, A.; Roffey, A.; Moth-Poulsen, K. Designing photoswitches for molecular solar thermal energy storage. Tetrahedron Lett. 2015, 56, 1457–1465. [Google Scholar] [CrossRef] [Green Version]
  21. Feng, Y.; Liu, H.; Luo, W.; Liu, E.; Zhao, N.; Yoshino, K.; Feng, W. Covalent functionalization of graphene by azobenzene with molecular hydrogen bonds for long-term solar thermal storage. Sci. Rep. 2013, 3, 3260. [Google Scholar] [CrossRef]
  22. Kurihara, S.; Nomiyama, S.; Nonaka, T. Photochemical control of the macrostructure of cholesteric liquid crystals by means of photoisomerization of chiral azobenzene molecules. Chem. Mater. 2001, 13, 1992–1997. [Google Scholar] [CrossRef]
  23. Kolpak, A.M.; Grossman, J.C. Hybrid chromophore/template nanostructures: A customizable platform material for solar energy storage and conversion. J. Chem. Phys. 2013, 138, 034303. [Google Scholar] [CrossRef]
  24. Wu, J.; Yang, Y.; Qu, Y.; Jia, L.; Zhang, Y.; Xu, X.; Chu, S.T.; Little, B.E.; Morandotti, R.; Jia, B.; et al. 2D Layered Graphene Oxide Films Integrated with Micro-Ring Resonators for Enhanced Nonlinear Optics. Small 2020, 16, 1906563. [Google Scholar] [CrossRef] [Green Version]
  25. Wu, J.; Yang, Y.; Qu, Y.; Xu, X.; Liang, Y.; Chu, S.T.; Little, B.E.; Morandotti, R.; Jia, B.; Moss, D.J. Graphene Oxide Waveguide and Micro-Ring Resonator Polarizers. Laser Photonics Rev. 2019, 13, 1900056. [Google Scholar] [CrossRef] [Green Version]
  26. Wu, S.; Butt, H.-J. Solar-Thermal Energy Conversion and Storage Using Photoresponsive Azobenzene-Containing Polymers. Macromol. Rapid Commun. 2020, 41, 1900413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Zhang, X.; Feng, Y.; Huang, D.; Li, Y.; Feng, W. Investigation of optical modulated conductance effects based on a graphene oxide–azobenzene hybrid. Carbon 2010, 48, 3236–3241. [Google Scholar] [CrossRef]
  28. Wang, Z.; Li, Z.-x.; Liu, Z. Photostimulated reversible attachment of gold nanoparticles on multiwalled carbon nanotubes. J. Phys. Chem. C 2009, 113, 3899–3902. [Google Scholar] [CrossRef]
  29. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  30. An, X.; Simmons, T.J.; Shah, R.; Wolfe, C.S.; Lewis, K.M.; Washington, M.; Nayak, S.K.; Talapatra, S.; Kar, S. Stable Aqueous Dispersions of Noncovalently Functionalized Graphene from Graphite and their Multifunctional High-Performance Applications. Nano Lett. 2010, 10, 4295–4301. [Google Scholar] [CrossRef]
  31. Englert, J.M.; Dotzer, C.; Yang, G.; Schmid, M.; Papp, C.; Gottfried, J.M.; Steinruck, H.; Spiecker, E.; Hauke, F.; Hirsch, A. Covalent bulk functionalization of graphene. Nat. Chem. 2011, 3, 279–286. [Google Scholar] [CrossRef] [PubMed]
  32. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef] [PubMed]
  33. Lim, D.J.; Marks, N.A.; Rowles, M.R. Universal Scherrer equation for graphene fragments. Carbon 2020, 162, 475–480. [Google Scholar] [CrossRef]
  34. Li, D.; Muller, M.B.; Gilje, S.; Kaner, R.B.; Wallace, G.G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 2008, 3, 101–105. [Google Scholar] [CrossRef]
  35. Min, M.; Bang, G.S.; Lee, H.; Yu, B. A photoswitchable methylene-spaced fluorinated aryl azobenzene monolayer grafted on silicon. Chem. Commun. 2010, 46, 5232–5234. [Google Scholar] [CrossRef]
  36. Gearba, R.I.; Lehmann, M.; Levin, J.; Ivanov, D.A.; Koch, M.H.J.; Barbera, J.; Debije, M.G.; Piris, J.; Geerts, Y. Tailoring discotic mesophases: Columnar order enforced with hydrogen bonds. Adv. Mater. 2003, 15, 1614–1618. [Google Scholar] [CrossRef]
  37. Zhang, X.; Feng, Y.; Lv, P.; Shen, Y.; Feng, W. Enhanced reversible photoswitching of azobenzene-functionalized graphene oxide hybrids. Langmuir 2010, 26, 18508–18511. [Google Scholar] [CrossRef]
  38. Zhang, B.; Zhang, Y.; Peng, C.; Yu, M.; Li, L.; Deng, B.; Hu, P.; Fan, C.; Li, J.; Huang, Q. Preparation of polymer decorated graphene oxide by γ-ray induced graft polymerization. Nanoscale 2012, 4, 1742–1748. [Google Scholar] [CrossRef]
  39. Yu, D.S.; Kuila, T.; Kim, N.H.; Khanra, P.; Lee, J.H. Effects of covalent surface modifications on the electrical and electrochemical properties of graphene using sodium 4-aminoazobenzene-4′-sulfonate. Carbon 2013, 54, 310–322. [Google Scholar] [CrossRef]
  40. Kim, M.; Safron, N.S.; Huang, C.; Arnold, M.S.; Gopalan, P. Light-driven reversible modulation of doping in graphene. Nano Lett. 2012, 12, 182–187. [Google Scholar] [CrossRef] [PubMed]
  41. Bandara, H.M.D.; Burdette, S.C. Photoisomerization in different classes of azobenzene. Chem. Soc. Rev. 2012, 41, 1809–1825. [Google Scholar] [CrossRef]
  42. Samanta, S.; Beharry, A.A.; Sadovski, O.; Mccormick, T.M.; Babalhavaeji, A.; Tropepe, V.; Woolley, G.A. Photoswitching azo compounds in vivo with red light. J. Am. Chem. Soc. 2013, 135, 9777–9784. [Google Scholar] [CrossRef] [PubMed]
  43. Kolpak, A.M.; Grossman, J.C. Azobenzene-Functionalized Carbon Nanotubes As High-Energy Density Solar Thermal Fuels. Nano Lett. 2011, 11, 3156–3162. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, G.; Amani, M.; Chaturvedi, A.; Tan, C.; Bullock, J.; Song, X.; Kim, H.; Lien, D.H.; Scott, M.C.; Zhang, H. Optical and electrical properties of two-dimensional palladium diselenide. Appl. Phys. Lett. 2019, 114, 253102. [Google Scholar] [CrossRef] [Green Version]
  45. Yang, Y.; Hughes, R.P.; Aprahamian, I. Visible Light Switching of a BF2-Coordinated Azo Compound. J. Am. Chem. Soc. 2012, 134, 15221–15224. [Google Scholar] [CrossRef]
  46. Siewertsen, R.; Neumann, H.; Buchheimstehn, B.; Herges, R.; Nather, C.; Renth, F.; Temps, F. Highly Efficient Reversible Z−E Photoisomerization of a Bridged Azobenzene with Visible Light through Resolved S1(nπ*) Absorption Bands. J. Am. Chem. Soc. 2009, 131, 15594–15595. [Google Scholar] [CrossRef]
  47. Liu, Y.; Grossman, J.C. Accelerating the Design of Solar Thermal Fuel Materials through High Throughput Simulations. Nano Lett. 2014, 14, 7046–7050. [Google Scholar] [CrossRef] [PubMed]
  48. Han, G.D.; Park, S.S.; Liu, Y.; Zhitomirsky, D.; Cho, E.; Dincă, M.; Grossman, J.C. Photon energy storage materials with high energy densities based on diacetylene–azobenzene derivatives. J. Mater. Chem. A 2016, 4, 16157–16165. [Google Scholar] [CrossRef]
  49. Zhitomirsky, D.; Cho, E.; Grossman, J.C. Solid-State Solar Thermal Fuels for Heat Release Applications. Adv. Energy Mater. 2016, 6, 1502006. [Google Scholar] [CrossRef]
Figure 1. (a,c) Low resolution TEM images of rGO and AzoF–rGO, (b,d) high resolution TEM images of rGO and AzoF-rGO with FFTs, and SEM images of (e) rGO and (f) AzoF–rGO.
Figure 1. (a,c) Low resolution TEM images of rGO and AzoF–rGO, (b,d) high resolution TEM images of rGO and AzoF-rGO with FFTs, and SEM images of (e) rGO and (f) AzoF–rGO.
Materials 14 01434 g001
Figure 2. (a) FT-IR spectra of rGO, AzoF and AzoF-rGO. (b) XPS spectra of rGO and AzoF-rGO. (c) F1s XPS spectra of AzoF-rGO. (d) TGA spectra of rGO, AzoF and AzoF-rGO.
Figure 2. (a) FT-IR spectra of rGO, AzoF and AzoF-rGO. (b) XPS spectra of rGO and AzoF-rGO. (c) F1s XPS spectra of AzoF-rGO. (d) TGA spectra of rGO, AzoF and AzoF-rGO.
Materials 14 01434 g002
Figure 3. UV–Vis absorption spectra of AzoF-rGO-3 (a) under UV irradiation, (b) in dark conditions, (c) under visible light irradiation, (d) reversion rates curves of AzoF-rGO in different environments.
Figure 3. UV–Vis absorption spectra of AzoF-rGO-3 (a) under UV irradiation, (b) in dark conditions, (c) under visible light irradiation, (d) reversion rates curves of AzoF-rGO in different environments.
Materials 14 01434 g003
Figure 4. Stability performance of (a) AzoF and (b) AzoF–rGO-3 for 50 times.
Figure 4. Stability performance of (a) AzoF and (b) AzoF–rGO-3 for 50 times.
Materials 14 01434 g004
Figure 5. DSC (differential scanning calorimetry) traces of (a) AzoF and (bd) AzoF-rGO after 1, 2 and 3-times reaction.
Figure 5. DSC (differential scanning calorimetry) traces of (a) AzoF and (bd) AzoF-rGO after 1, 2 and 3-times reaction.
Materials 14 01434 g005
Figure 6. Power density of AzoF and AzoF-rGO after 1, 2 and 3-times reaction.
Figure 6. Power density of AzoF and AzoF-rGO after 1, 2 and 3-times reaction.
Materials 14 01434 g006
Table 1. Ad of AzoF on rGO.
Table 1. Ad of AzoF on rGO.
TGAXPS
Reaction TimesDg (%) aAdElement Content (%)Ad
CFO
AzoF-rGO-143.411:40.177.424.1315.711:40.2
AzoF-rGO-252.951:27.374.135.0917.391:27.7
AzoF-rGO-365.731:16.071.076.6417.901:16.1
aDg is the average weight percentage of AzoF in AzoF-rGO at 600 °C, 700 °C and 800 °C.
Table 2. Performance of different photothermal conversion materials.
Table 2. Performance of different photothermal conversion materials.
Photothermal Conversion MaterialEnergy Density
(kJ mol−1)
Power Density
(W mol−1)
Half-Life (h)Ref.
Azo-diacetylene polymer176.21289.527.8[48]
Azo-SWCNT complex92.0457.10.5[7]
Azo-PCM complex79.3[15]
Azo-alkyl polymer89.0148.655[49]
AzoF-rGO-3 complex367.7848.687.7This paper
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, X.; Li, S.; Zhang, J.; Wang, X.; Wang, Y.; Zhao, J. A Low-Temperature Heat Output Photoactive Material-Based High-Performance Thermal Energy Storage Closed System. Materials 2021, 14, 1434. https://doi.org/10.3390/ma14061434

AMA Style

Yang X, Li S, Zhang J, Wang X, Wang Y, Zhao J. A Low-Temperature Heat Output Photoactive Material-Based High-Performance Thermal Energy Storage Closed System. Materials. 2021; 14(6):1434. https://doi.org/10.3390/ma14061434

Chicago/Turabian Style

Yang, Xiangyu, Shijie Li, Jin Zhang, Xiaomin Wang, Yongzhen Wang, and Jianguo Zhao. 2021. "A Low-Temperature Heat Output Photoactive Material-Based High-Performance Thermal Energy Storage Closed System" Materials 14, no. 6: 1434. https://doi.org/10.3390/ma14061434

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop