Next Article in Journal
Pencil Graphite Electrocatalytic Sensors Modified by Pyrene Coated Reduced Graphene Oxide Decorated with Molybdenum Disulfide Nanoroses for Hydrazine and 4-Nitrophenol Detection in Real Water Samples
Previous Article in Journal
Evaluating the Anti-Melanoma Effects and Toxicity of Cinnamaldehyde Analogues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Doped Graphene Quantum Dot-Passivated δ-Phase CsPbI3: A Water-Stable Photocatalytic Adjuvant to Degrade Rhodamine B

College of Science & Key Laboratory of Low-Dimensional Structural Physics and Application, Education Department of Guangxi Zhuang Autonomous Region, Guilin University of Technology, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(21), 7310; https://doi.org/10.3390/molecules28217310
Submission received: 24 September 2023 / Revised: 22 October 2023 / Accepted: 26 October 2023 / Published: 28 October 2023
(This article belongs to the Special Issue Novel Materials for Ion Batteries)

Abstract

:
Inorganic halide perovskite CsPbI3 is highly promising in the photocatalytic field for its strong absorption of UV and visible light. Among the crystal phases of CsPbI3, the δ-phase as the most aqueous stability; however, directly using it in water is still not applicable, thus limiting its dye photodegradation applications in aqueous solutions. Via adopting nitrogen-doped graphene quantum dots (NGQDs) as surfactants to prepare δ-phase CsPbI3 nanocrystals, we obtained a water-stable material, NGQDs-CsPbI3. Such a material can be well dispersed in water for a month without obvious deterioration. High-resolution transmission electron microscopy and X-ray diffractometer characterizations showed that NGQDs-CsPbI3 is also a δ-phase CsPbI3 after NGQD coating. The ultraviolet-visible absorption spectra indicated that compared to δ-CsPbI3, NGQDs-CsPbI3 has an obvious absorption enhancement of visible light, especially near the wavelength around 521 nm. The good dispersity and improved visible-light absorption of NGQDs-CsPbI3 benefit their aqueous photocatalytic applications. NGQDs-CsPbI3 alone can photodegrade 67% rhodamine B (RhB) in water, while after compositing with TiO2, NGQDs-CsPbI3/TiO2 exhibits excellent visible-light photocatalytic ability, namely, it photodegraded 96% RhB in 4 h. The strong absorption of NGQDs-CsPbI3 in the visible region and effective transfer of photogenerated carriers from NGQDs-CsPbI3 to TiO2 play the key roles in dye photodegradation. We highlight NGQDs-CsPbI3 as a water-stable halide perovskite material and effective photocatalytic adjuvant.

1. Introduction

In recent years, with the rapid development of the economy and technology, organic dyes have been widely used and discharged into industrial wastewater. Especially, rhodamine B (RhB), which is widely used for dyeing fabrics, paints, acrylics, and biological products, has become abundant in wastewater and highly toxic to organisms [1]. Previous studies have shown that RhB can induce growth retardation and liver damage, erythrocyte hemolysis, and suppression of the immune response in isolated spleen cells [2]. Other studies have suggested that RhB is mutagenic and carcinogenic [3] and could produce local sarcomas [2]. Therefore, it is urgent to develop economic and effective ways to remove RhB in wastewater. Producing highly oxidative active species is the basic technique for removing organic pollutants. Photocatalysis is a feasible and economic way to produce oxidative species and has been widely studied [4,5,6]. Fujishima and Honda first reported the photocatalytic performance of TiO2 [7], and since then many scholars have investigated its photodegradation effect [8,9]. However, its wide band gap of 3.2 eV determines that only UV light can be absorbed, and visible light, the main component of sunlight, cannot be well captured by TiO2 [10]. To obtain a visible-light photocatalyst, one way is to choose a photocatalytic material with a narrower band gap. Typically, graphitic C3N4, with a band gap of ~2.71 eV, can absorb visible light and has been proven effective in visible-light photodegradation [11], but the high carrier recombination rate remarkably restrains its photocatalytic performance [12]. The other way is compositing a visible-light responsive photocatalytic adjuvant with TiO2, which turns visible light into photogenerated carriers and effectively transfers them to TiO2, and then charged oxidative species are produced to degrade organic dyes. Carbon-based materials are universal adjuvants due to their widely tunable bandgaps, ease of composition, good aqueous dispersibility, and availability of carrier separation and transfer [13,14,15,16]. However, carbon-based materials suffer from weak visible-light absorption ability because their absorption mainly comes from the energy transition from σ or π to π* orbitals [17]. Searching for an alternative adjuvant with strong absorption in the visible region is beneficial to improving the visible-light photocatalytic performance of TiO2.
Halide perovskite, the rising star in optoelectronics, has been successfully applied in photocatalytic fields such as carbon dioxide reduction [18], water splitting [19], and dye removal [20]. Its strong visible-light acquisition ability plays the key role in photocatalysis. However, poor water stability generally is the Archilles heel of halide perovskite, so such experiments have to be conducted in organic solvents or after water-resistant coating [18,19,20]. With the invasion of water, the perovskite structure tends to stretch, distort, and tilt, turning α-, β-, and γ-phases into δ-phase [21]. δ-phase halide perovskite is generally deemed as a waste phase because it no longer exhibits typical perovskite characteristics, for example, strong photoluminescence (PL) and high quantum yield (QY) [22,23]. Also, because of this, δ-phase perovskite is the most water stable, and the low PLQY means very few photogenerated carriers undergo direct recombination. Both of these properties are very suitable for aqueous photocatalysis.
Herein, we studied the photodegradation performance of δ-CsPbI3 nanocrystals on RhB. The δ-CsPbI3 nanocrystals synthesized by the conventional method using oleic acid and oleylamine as surfactants tend to aggregate and cannot be well dispersed in water, so it is hard to perform photodegradation experiments. By using nitrogen-doped graphene quantum dots (NGQDs) as surfactants to synthesize δ-phase CsPbI3 [24], we obtained NGQDs-CsPbI3, which could be perfectly dispersed and stably stored in water for a month. After compositing with TiO2, under visible light, the photogenerated electrons in NGQDs-CsPbI3 transfer to TiO2, producing the oxidant radicals •O2 and •OH, and the photogenerated holes left in NGQDs-CsPbI3 combine with H2O/OH to produce •OH, finally oxidizing RhB into mineralization products. NGQDs-CsPbI3/TiO2 exhibits excellent visible-light photocatalytic ability, which could photodegrade 96% RhB in 4 h. As a water-stable material, NGQDs-CsPbI3 shows bright prospects in photocatalysis, and it also opens the door to resurrecting the potential of δ-phase halide perovskites.

2. Results

Typical TEM images of the synthesized samples are shown in Figure 1. The NGQDs are sized 1–10 nm [Figure S1a], with 65.09, 24.05, and 10.76 at.% of C, O, and N, respectively [Figure S1b]. Figure 1a shows a single NGQD of ~5 nm with a lattice stripe distance of ~0.22 nm, which is similar to that of graphite 11 2 ¯ 0 facets [25]. The δ-phase CsPbI3 nanocrystals are sized 5–20 nm [Figure S2a]. A typical 5 nm nanocrystal is shown in Figure 1b, and the lattice stripes are clearly resolved with interplanar spacing of 0.33 nm, corresponding to (212) planes of orthorhombic δ-CsPbI3 [26]. The borders of NGQDs-CsPbI3 crystals are extremely irregular [Figure 1c], and their sizes are observably enlarged, obviously ascribed to the outcome of adopting NGQDs as surfactants. Figure 1d shows a large-size NGQDs-CsPbI3 microcrystal, and its atomic distribution can be clearly seen with interplanar spacing of 0.35 nm. Figure S2c,d shows two other NGQDs-CsPbI3 nano- or microcrystals, for which different crystal spacings are exhibited. One possibility is that these spacings correspond to different crystal planes of δ-CsPbI3, and another is that the crystal structures of NGQDs-CsPbI3 are miscellaneous since nonstoichiometric NGQDs might bring about different strains to build δ-CsPbI3 with different crystal constants [27]. Among the large particles of TiO2, the NGQDs-CsPbI3 crystals can also be resolved [see the yellow circle in Figure 1e]. From the crystal plane distance of 0.35 nm, one can know it is a NGQDs-CsPbI3 nanocrystal [Figure 1f].
The EDS elemental mapping of an NGQDs-CsPbI3 microcrystal is shown in Figure 2. Typical elements of NGQDs-CsPbI3, such as Cs, Pb, I, C, and N, can be found uniformly distributed. From Figure S3, one can know the atomic ratio of Cs:Pb:I ≈ 1:1:3, according to the formula of halide perovskite. The atomic ratio of Cs:C is ~1:8, and it is must be stressed that the NGQD portion in NGQDs-CsPbI3 may be overrated since NGQDs are used as surfactants and attached on the CsPbI3 surface.
The XRD patterns are shown in Figure 3a. The prominent peaks of δ-CsPbI3 are (002), (102), (200), (201), (111), (112), (210), (113), (212), and (302) at 10.27°, 13.43°, 17.42°, 18.20°, 21.41°, 23.27°, 25.70°, 26.03°, and 27.55°, respectively, which are totally in accordance with the standard PDF25-0744 of orthorhombic δ-phase CsPbI3 [28]. NGQDs-CsPbI3 exhibits similar characteristics, and in its XRD plot, the typical (002) peak of NGQDs is found at around 26° and is not well resolved [29]. Although some peaks are inconspicuous, NGQDs-CsPbI3 can still be reasonably seen as orthorhombic δ-phase CsPbI3 with NGQDs on the surface. The XRD pattern of NGQDs-CsPbI3/TiO2 [Figure 3b] demonstrates no new peaks aside from those of NGQDs-CsPbI3 and TiO2, meaning that the two parts are merely physically composited.
The aqueous stability of the photocatalyst is a key factor on deciding whether it can be used in water to photodegrade organic dyes. In Figure 4a, one can find that the newly synthesized δ-CsPbI3 nanocrystals can be well dispersed in water; however, shortly after preparation, they are inclined to aggregate and hard to redisperse even with stirring, so δ-CsPbI3 nanocrystals are hardly used as aqueous photocatalysts. On the contrary, NGQDs-CsPbI3 crystals could be well dispersed for a month, which is obviously highly related to the good dispersibility of NGQDs in water [30]. Here, we highlight the effect of nitrogen doping in enhancing dispersion, since we replaced NGQDs with graphene quantum dots (GQDs) to prepare GQDs-CsPbI3 under the same conditions and found that, totally like δ-CsPbI3, GQDs-CsPbI3 tended to aggregate quickly. In Figure 4b, after 30 days, the PL shapes and intensities of NGQDs-CsPbI3 nanocrystals change very slightly, demonstrating the water stability of NGQDs-CsPbI3. It can also be found that the PL peak of NGQDs-CsPbI3 can be deconvoluted into two parts: one is ascribed to NGQDs centered at 461 nm, and the other is ascribed to δ-CsPbI3 centered at 521 nm. The δ-CsPbI3 part demonstrates the typical green light of δ-phase iodide perovskite, corresponding to self-trapped exciton emission [31]. In Figure 4c, via Tauc plotting, one can find that the bandgap of NGQDs-CsPbI3 (3.09 eV) is a bit larger than that of δ-CsPbI3 (2.8 eV); however, in the visible region, especially around 521 nm, the absorption of NGQDs-CsPbI3 is obviously stronger. Such absorption is ascribed to the transition from the ground states to trap states, namely, passivation through NGQDs can enhance the trapping exciton absorption of δ-CsPbI3. The improvement in visible-light absorption undoubtedly benefits the photocatalytic ability of NGQDs-CsPbI3.
Figure 4d shows the PLs of TiO2 and NGQDs-CsPbI3 before and after compositing. In the PL spectrum of NGQDs-CsPbI3/TiO2, the subpeak ascribed to δ-CsPbI3 (centered at 521 nm) is almost quenched, and the PL of TiO2 (centered at 385 nm) is also invisible. NGQDs-CsPbI3/TiO2 emits light with a single peak centered at ~440 nm, representing the characteristics of NGQDs, since it only exhibits a slight blue-shift compared to NGQDs alone (~460 nm). Therefore, we can speculate that the photogenerated electrons or holes in TiO2 and CsPbI3 are effectively transferred, and some of them recombine on NGQDs, namely on the coating surface of NGQDs-CsPbI3 crystals. In Figure 4e, by tuning the NGQD mass ratio in NGQDs-CsPbI3 (the NGQD mass is 0.5, 0.7, and 0.9 mg in NGQDs-1-CsPbI3, NGQDs-CsPbI3, and NGQDs-2-CsPbI3, respectively, and the CsPbI3 mass is 72 mg), it is found that the PL positions remain invariable, merely along with sightly increased PL intensity due to the increase in NGQD weight. In a word, by compositing NGQDs-CsPbI3 with TiO2, the photoinduced carriers in both can be effectively separated and transferred, which is highly beneficial to photocatalytic applications [32].
Taking 100 mL RhB water solution (10 mg/L) as a reference, we studied the visible-light (λ > 420 nm) photodegradation activities of NGQDs-CsPbI3/TiO2 (72.7 mg/250 mg). The effects of TiO2 (250 mg), NGQDs (7 mg), NGQDs/TiO2 (7 mg/250 mg), and NGQDs-CsPbI3 (72.7 mg) are also supplied for comparison. The photodegradation results are shown in Figure 5a, where C0 and C are the initial and real-time concentrations of RhB, respectively, and C/C0 is determined by the absorbance of RhB at 554 nm. Due to the bad dispersibility of δ-CsPbI3, it was hard to assess its photodegradation ability, so the related results are not shown here.
Obviously, the visible-light photocatalytic ability of TiO2 or NGQDs alone is negligible, and the RhB photodegradation ratio is enhanced after compositing [30]. It is worth noting that NGQDs-CsPbI3 alone displays decent photocatalytic activity, degrading 67% RhB in 4 h (see Section 4.4 for the calculation of photodegradation efficiency). The PLQY of NGQDs-CsPbI3 is ~3% (Table S1) and together with its broad absorption range [Figure 4c], many indirect recombined photogenerated carriers can effectively participate in the photodegradation process. To further improve the photocatalytic activity, those photocarriers must be transferred to reduce the possibility of direct recombination. By compositing NGQDs-CsPbI3 with TiO2, RhB can be nearly completely photodegraded (96%) after 4 h. Using total organic carbon (TOC) as a reference to measure the mineralization rate of RhB, similar photodegradation efficiency can be obtained, which is 94% in 4 h (Figure S4). Since the difference in photodegradation efficiency obtained by these two references (RhB absorbance and TOC) is small, all further discussion about the photodegradation of RhB is based on the former reference. The aforementioned discussion about Figure 4d shows that photocarriers are effectively transferred between TiO2 and NGQDs-CsPbI3 and TiO2 is almost entirely inactive to visible light; hence, effective photocarrier transfer from NGQDs-CsPbI3 to TiO2 is the reason why the photodegradation activity can be improved. The corresponding photodegradation kinetics were fitted using the first-order reaction equation:
l n C C 0 = k t
where k is the photocatalytic efficiency [33]. The k value of NGQDs-CsPbI3/TiO2 is 0.85, almost 3 times that of NGQDs-CsPbI3 alone [0.26, Figure 5b]. In order to further evaluate the usefulness of NGQDs-CsPbI3/TiO2 nanocrystals as photocatalytic materials, cyclic experiments of RhB photodegradation were performed [Figure 5c]. Regrettably, a slight decrease in photodegradation activity appears in the 3rd to 5th cycles, with 92%, 79%, and 77%, respectively. This is due to the mass loss of NGQDs-CsPbI3/TiO2, especially NGQDs-CsPbI3 nanocrystals, since it is difficult to completely collect nano-size particles in water through high-speed centrifugation. In Figure S5, one can find that using different ratios of NGQDs as surfactants to synthesize NGQDs-CsPbI3 changes the photodegradation ability of NGQDs-CsPbI3/TiO2, and the optimal ratio is adopted in our experiments.
Generally, photogenerated •O2, holes (h+), and •OH play important roles in the degradation of organic dyes [34]. In order to identify which free radicals are dominant in photodegradation by NGQDs-CsPbI3/TiO2, radical capture experiments were conducted. As shown in Figure 5d, three scavengers, benzoquinone (BQ, 0.01 g/L), disodium ethylenediaminetetraacetate (EDTA-2Na, 0.02 g/L), and isopropyl alcohol (IPA, 0.02 g/L), were used in this study to capture the •O2, h+, and •OH radicals, respectively [35]. The RhB degradation rates of NGQDs-CsPbI3/TiO2 nanocomposites were 21%, 33%, and 47% in the presence of EDTA-2Na, BQ, and IPA, respectively. The results showed that in our photocatalytic process, these three active components, •O2, •OH, and h+, are all massively produced and participate in dye oxidation, and the roles of •O2 and h+ are a bit more important than that of •OH. Regrettably, several capture agents (including AgNO3, K2Cr2O7, and KBrO3) were used to capture e, and it was found that RhB was rapidly adsorbed and it was difficult to assess the photodegradation ability after e capture.
In addition, Table 1 lists a comparison of relevant photocatalysts that have been used in several studies to degrade contaminants. Under visible light, the photodegradation ability of TiO2 is negligible. Most halide perovskites can photodegrade dyes effectively, and these results show a good foreground of photocatalytic applications of perovskites. However, such experiments have been conducted in organic solution to ensure the stability of perovskites. As is known, the removal of organic contamination in wastewater is the first thing to be resolved; hence, to a certain degree, developing water-stable photocatalysts is more important than enhancing the photocatalytic efficiency alone. Obviously, NGQDs-CsPbI3/TiO2 is a promising photocatalyst due to its feasibility of direct use in water, although there is much room for improvement in the photocatalytic efficiency.

3. Discussion

Figure 6 shows a schematic diagram of the RhB photodegradation process of NGQDs-CsPbI3/TiO2. Using the vacuum level as a reference, the UPS measurement [inset of Figure 6] can define the valence band maximum (VBM) by EVBM = −[21.22 − (Ecutoff − Eonset)]. For NGQDs-CsPbI3, the Ecutoff is 16.53 eV and the Eonset is 1.59 eV, so the EVBM is −6.12 eV. Considering the band gap is 3.09 eV, we know that the maximum conduction band (CBM) is −3.03 eV. The emissive center is located at 521 nm, implying that the energy difference between the trap states and VBM is 2.38 eV, namely, the energy level of the trap states is −3.74 eV. Conventionally, we express energy levels using normal hydrogen electrode (NHE) as a reference [E(NHE) = −4.5 − E(vacuum), which represents the relationship between NHE and the vacuum energy levels], and the energy levels of CBM, the trap states, and VBM are −1.47, −0.76, and 1.62 eV, respectively. The photocatalytic mechanism of NGQDs-CsPbI3/TiO2 can be reasonably inferred according to the following expressions (2)–(8):
N G Q D s C s P b I 3 / T i O 2 + d y e s d y e s
h ν o n   N G Q D s C s P b I 3 h + + e
where *dye denotes the dye molecules activated by NGQDs-CsPbI3/TiO2. When NGQDs-CsPbI3/TiO2 is exposed to visible light, electrons and holes are only photogenerated in NGQDs-CsPbI3 since TiO2 is inactive under visible light. It also must be stressed that the electrons in NGQDs-CsPbI3 can only be excited from VBM to the trap states since the photon energy (λ > 420 nm) cannot compensate for the bandgap of NGQDs-CsPbI3 (3.09 eV). The primary function process of •OH radicals is shown as the following expressions (4)–(6):
H 2 O     H + + O H
h + + O H O H
O H + d y e s m i n e r a l i z a t i o n   p r o d u c t s
As shown in Figure 6, the photogenerated holes in NGQDs-CsPbI3 combine with OH generated by water to form •OH, and then •OH reacts with *dyes to produce mineralization products. Generally speaking, oxidizing OH into •OH requires the holes to have a high potential. For example, in TiO2 [39], Bi-doped LaFeO3 [40], and g-C3N4 [41], the hole potentials are 1.83, 1.97, and 1.84 V, respectively. However, in halide perovskite, this potential can decrease to as low as 1.1 V [36]. The VBM of NGQDs-CsPbI3 is 1.62 eV, which might be enough for holes to generate •OH radicals in such a perovskite material. Moreover, the broad PL and wide absorption of NGQDs-CsPbI3 indicate that the excited electrons and holes are distributed in a wide energy range, which means the potentials of many holes are higher than 1.62 V, further boosting the oxidation of OH into •OH to degrade pollutants. The photodegradation process of •O2 radicals can also be seen in Figure 6, as follows:
O 2 + e O 2
O 2 + d y e s m i n e r a l i z a t i o n   p r o d u c t s
The electrons produced by photoexcitation are transferred from the trap states of NGQDs-CsPbI3 to the CBM of TiO2, combine with the adsorbed O2 to form •O2, and then •O2 reacts with *dyes to form mineralized compounds. Certainly, some •O2 can be captured by H2O2 to form •OH, and again, •OH reacts with *dyes to form mineralized products.

4. Materials and Methods

4.1. Materials

Citric acid (CA, anhydrous), urea (99.5%), cesium iodide (CsI, 99.9%), oleylamine (OLA, 96%), dimethylformamide (DMF, 99.8%), and RhB (98%) were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). Lead iodide (PbI2, 99.998%) was purchased from Alfa. Oleic acid (OA, 90%), P25 TiO2 (80% anatase and 20% rutile), isopropyl alcohol (IPA), p-benzoquinone (BQ), and disodium EDTA-2Na were purchased from Xilong Chemical Co., Ltd. (Shantou, China). All chemicals were used directly without further purification.

4.2. Sample Synthesis

4.2.1. Synthesis of NGQDs

First, 0.53 g CA and 0.6 g urea were dissolved in 12 mL deionized water and stirred to form a clarified solution. The solution was then transferred to a 50 mL Teflon autoclave. The sealed autoclave was heated to 160 °C in an oven and maintained for 8 h. The final product was collected by adding ethanol to the solution and centrifuging at 5000 rpm for 5 min. Finally, the sediment was dried at 60 °C to obtain NGQDs.

4.2.2. Synthesis of δ-CsPbI3

First, 0.26 g CsI and 0.46 g PbI2 were dissolved in 2 mL DMF, and the solution was heated to 50 °C and held at this temperature for 20 min. Subsequently, 0.4 mL OA and 0.2 mL OLA were added to stabilize the precursor solution, and then the precursor solution was heated to 90 °C and held at this temperature for 30 min. Finally, 0.2 mL precursor solution was quickly added to 4 mL deionized water (under vigorous stirring); after drying, δ-CsPbI3 nanocrystals were obtained.

4.2.3. Synthesis of NGQDs-CsPbI3

The synthesis process of NGQDs-CsPbI3 was similar to that of δ-CsPbI3, except that the organic ligands OA and OLA were replaced by NGQDs. First, 0.26 g CsI and 0.46 g PbI2 were dissolved in 2 mL DMF. Next, 7 mg NGQDs was added to obtain a precursor solution. Then, 0.2 mL precursor solution was rapidly added to 4 mL of deionized water (under vigorous stirring) to immediately generate NGQDs-CsPbI3 crystals. For comparison, the initial mass ratio of NGQDs was tuned to synthesize NGQDs-CsPbI3, and these were named as NGQDs-1-CsPbI3 and NGQDs-2-CsPbI3 with initial NGQD masses of 5 and 9 mg, respectively.

4.2.4. Synthesis of NGQDs-CsPbI3/TiO2

First, 250 mg TiO2 was added into the above NGQDs-CsPbI3 solution (including 72.7 mg NGQDs-CsPbI3, and the mass of NGQDs is 0.7 mg) and uniformly mixed with stirring; after drying, NGQDs-CsPbI3/TiO2 composite was obtained.

4.3. Characterization

The crystal structures were characterized using a Miniflex-600 X-ray diffractometer (XRD, JEOL, Tokyo, Japan). The morphologies were assessed using a field emission transmission electron microscope (TEM, JEM2100F, Tokyo, Japan) equipped with a selected area electron diffractometer (SAED) and a scanning electron microscope (SEM, TESCAN MIRA LMS, Brno, Czech Republic) equipped with an energy dispersive spectroscope (EDS). Surface analyses were carried out with an Escalab-250XI X-ray photoelectron spectrometer (XPS) from Thermo Fisher Scientific (Waltham, MA, USA). Ultraviolet-visible (UV-vis) absorption spectra were obtained using a PerkinElmer Lambda 750. Photoluminescence (PL) spectra were determined using an Edinburgh FL/FS900 carry Eclipse (Cheadle, UK). UV photoelectron spectroscopy (UPS) measurements were performed on a photoelectron spectrometer (ESCALAB 250Xi) with a He I source of 21.22 eV.

4.4. Photodegradation Test

The photocatalytic activities were investigated by photodegrading RhB (100 mL, 10 mg/L) under a xenon lamp (PLS-SXE 300, 300 W, λ > 420 nm). The adsorption–desorption equilibrium of RhB and photocatalysts was achieved after being stirred in the dark for 0.5 h. Every 0.5 h, 4 mL solution was taken out to determine its RhB concentration using a Lambda 950 UV-Vis spectrophotometer. The recycle photodegradation experiments were conducted by repeatedly collecting the photocatalysts via centrifugation and drying.
The concentration of RhB was measured by UV-vis spectrophotometry and the degradation efficiency of RhB was calculated as:
d e g r a d a t i o n   r a t i o   ( % ) = C 0 C C 0 × 100 = A 0 A A 0 × 100
where C0 and C are the initial and real-time concentrations of RhB, respectively; and A0 and A are the initial and real-time absorbance of RhB at 554 nm, respectively. The mineralization rate of RhB solutions was measured using a TOC analyzer (Shimadzu, TOC-L CPN, Tsushima, Japan). The photodegradation ratio was determined by the following formula:
d e g r a d a t i o n   r a t i o   ( % ) = T O C 0 T O C T O C 0 × 100
where TOC0 and TOC are respectively the initial and real-time TOC of RhB solution.

5. Conclusions

Water-stable NGQDs-CsPbI3 halide perovskite was successfully prepared by the thermal injection method at room temperature. It exhibits typical characteristics of δ-phase CsPbI3, namely strong absorption in the visible region and low PL QY. Due to the coating and passivation through NGQDs, such a material can be well dispersed and is stable in water for a month, hence it can effectively photodegrade RhB. After compositing with TiO2, the photodegradation ability is further improved because the photogenerated e-h pairs can be effectively separated and transferred. We highlight NGQDs-CsPbI3 as a water-stable perovskite with great potential in the photocatalytic field.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28217310/s1, Figure S1. (a) TEM image of NGQDs. (b) Statistical size distribution of the prepared NGQDs. (c) XPS full spectrum of NGQDs. (d) High-resolution of N 1s and C 1s spectra of NGQDs. Figure S2. (a) TEM images of δ-phase CsPbI3 nanocrystals. (b) Statistical size distribution of the prepared δ-phase CsPbI3. (c) and (d) High-resolution TEM images of NGQDs-CsPbI3 with different space stripes. Figure S3. (a) EDS mapping and (b) the atomic proportion of NGQDs-CsPbI3. Figure S4. Comparison of the photodegradation ratios of RhB obtained by the absorbance (Abs) and total organic carbon (TOC) analysis of RhB solution. Figure S5. The RhB photodegradation activities of different NGQDs-CsPbI3/TiO2 samples by tuning the initial NGQDs mass. The NGQDs mass is 0.5, 0.7 and 0.9 mg in NGQDs-1-CsPbI3, NGQDs-CsPbI3 and NGQDs-2-CsPbI3, respectively, and the CsPbI3 part is 72 mg. The mass of TiO2 is 250 mg. Table S1. Quantum yields of δ-CsPbI3, NGQDs, NGQDs-CsPbI3 and NGQDs-CsPbI3/TiO2 using quinine sulfate as a reference.

Author Contributions

Formal analysis, Writing—original draft preparation, Data curation, Y.G.; Investigation, Y.G., X.D. and F.H.; Conceptualization, Software, J.W.; Methodology, X.D. and F.H.; Supervision, T.T. and M.L.; Visualization, Y.G. and T.T.; Resources, Validation, Writing—review and editing, Project administration, T.T.; Funding acquisition, T.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Guangxi Zhuang Autonomous Region of China (2018GXNSFAA050014) and the Foundation of Guilin University of Technology (GLUTQD2002023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data can be made available upon reasonable request.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Sample Availability

Not applicable.

References

  1. Xu, D.; Ma, H. Degradation of rhodamine B in water by ultrasound-assisted TiO2 photocatalysis. J. Clean. Prod. 2021, 313, 127758. [Google Scholar]
  2. Elliott, G.S.; Mason, R.W.; Edwards, I.R. Studies on the pharmacokinetics and mutagenic potential of rhodamine B. J. Toxicol.-Clin. Toxic. 1990, 28, 45–59. [Google Scholar]
  3. Gupta, V.K.; Suhas; Ali, I.; Saini, V.K. Removal of rhodamine B, fast green, and methylene blue from wastewater using red mud, an aluminum industry waste. Ind. Eng. Chem. Res. 2004, 43, 1740–1747. [Google Scholar]
  4. Liu, Y.; Cheng, M.; Liu, Z.; Zeng, G.; Zhong, H.; Chen, M.; Zhou, C.; Xiong, W.; Shao, B.; Song, B. Heterogeneous fenton-like catalyst for treatment of rhamnolipid-solubilized hexadecane wastewater. Chemosphere 2019, 236, 124387. [Google Scholar]
  5. Liu, Y.; Huang, D.; Cheng, M.; Liu, Z.; Lai, C.; Zhang, C.; Zhou, C.; Xiong, W.; Qin, L.; Shao, B.; et al. Metal sulfide/MOF-based composites as visible-light-driven photocatalysts for enhanced hydrogen production from water splitting. Coord. Chem. Rev. 2020, 409, 213220. [Google Scholar]
  6. Xiao, S.; Cheng, M.; Zhong, H.; Liu, Z.F.; Liu, Y.; Yang, X.; Liang, Q.H. Iron-mediated activation of persulfate and peroxymonosulfate in both homogeneous and heterogeneous ways: A review. Chem. Eng. J. 2020, 384, 123265. [Google Scholar]
  7. Emin, S.; Singh, S.P.; Han, L.; Satoh, N.; Islam, A. Colloidal quantum dot solar cells. Nat. Energy 2011, 85, 1264–1282. [Google Scholar]
  8. Khalid, N.R.; Mazia, U.; Tahir, M.B.; Niaz, N.A.; Javid, M.A. Photocatalytic degradation of RhB from an aqueous solution using Ag3PO4/N-TiO2 heterostructure. J. Mol. Liq. 2020, 313, 113522. [Google Scholar]
  9. Lin, X.; Li, Y.X. Preparation of TiO2/Ag BMIM Cl composites and their visible light photocatalytic properties for the degradation of rhodamine B. Catalysts 2021, 11, 661. [Google Scholar] [CrossRef]
  10. Todescato, F.; Fortunati, I.; Gardin, S.; Garbin, E.; Collini, E.; Bozio, R.; Jasieniak, J.J.; Della Giustina, G.; Brusatin, G.; Toffanin, S.; et al. Soft-lithographed up-converted distributed feedback visible lasers based on CdSe-CdZnS-ZnS quantum dots. Adv. Funct. Mater. 2012, 22, 337–344. [Google Scholar] [CrossRef]
  11. Guo, Z.; Ni, S.; Wu, H.; Wen, J.; Li, X.; Tang, T.; Li, M.; Liu, M. Designing nitrogen and phosphorus co-doped graphene quantum dots/g-C3N4 heterojunction composites to enhance visible and ultraviolet photocatalytic activity. Appl. Surf. Sci. 2021, 548, 149211. [Google Scholar]
  12. Liu, Y.; Zeng, X.; Hu, X.; Hu, J.; Wang, Z.; Yin, Y.; Sun, C.; Zhang, X. Two-dimensional g-C3N4/TiO2 nanocomposites as vertical Z-scheme heterojunction for improved photocatalytic water disinfection. Catal. Today 2019, 335, 243–251. [Google Scholar]
  13. Zhang, L.; Zhang, Q.; Xie, H.; Guo, J.; Lyu, H.; Li, Y.; Sun, Z.; Wang, H.; Guo, Z. Electrospun titania nanofibers segregated by graphene oxide for improved visible light photocatalysis. Appl. Catal. B Environ. 2017, 201, 470–478. [Google Scholar]
  14. Li, K.; Xiong, J.; Chen, T.; Yan, L.; Dai, Y.; Song, D.; Lv, Y.; Zeng, Z. Preparation of graphene/TiO2 composites by nonionic surfactant strategy and their simulated sunlight and visible light photocatalytic activity towards representative aqueous POPs degradation. J. Hazard. Mater. 2013, 250–251, 19–28. [Google Scholar]
  15. Ioan-Alexandru, B.; John, B.; Kiem, G.N.; Tobias, H.; Muhammad Tariq, S.; Stuart, A.J.T.; Alistair, R.; David, J.M.; Nicholas, P.P.; Sabina Alexandra, N.; et al. Outstanding visible light photocatalysis using nano-TiO2 hybrids with nitrogen-doped carbon quantum dots and/or reduced graphene oxide. J. Mater. Chem. A 2023, 11, 9791–9806. [Google Scholar]
  16. Guo, Z.; Wu, H.; Li, M.; Tang, T.; Wen, J.; Li, X. Phosphorus-doped graphene quantum dots loaded on TiO2 for enhanced photodegradation. Appl. Surf. Sci. 2020, 526, 146724. [Google Scholar]
  17. Li, M.; Wu, W.; Ren, W.; Cheng, H.; Tang, N.; Zhong, W.; Du, Y. Synthesis and upconversion luminescence of N-doped graphene quantum dots. Appl. Phys. Lett. 2012, 101, 103107. [Google Scholar]
  18. Pan, A.; Ma, X.; Huang, S.; Wu, Y.; Jia, M.; Shi, Y.; Liu, Y.; Wangyang, P.; He, L.; Liu, Y. CsPbBr3 perovskite nanocrystal grown on MXene nanosheets for enhanced photoelectric detection and photocatalytic CO2 reduction. J. Phys. Chem. Lett. 2019, 10, 6590–6597. [Google Scholar]
  19. Nam, S.; Mai, C.T.K.; Oh, I. Ultrastable photoelectrodes for solar water splitting based on organic metal halide perovskite fabricated by lift-off process. ACS Appl. Mater. Interfaces 2018, 10, 14659–14664. [Google Scholar]
  20. Li, K.; Li, S.; Zhang, W.; Shi, Z.; Wu, D.; Chen, X.; Lin, P.; Tian, Y.; Li, X. Highly-efficient and stable photocatalytic activity of lead-free Cs2AgInCl6 double perovskite for organic pollutant degradation. J. Colloid Interface Sci. 2021, 596, 376–383. [Google Scholar]
  21. Huang, Z.; Qin, H.; Wen, J.; Jiang, L.; Hu, G.; Li, M.; Chen, J.; Liu, F.; Tao, T. Observation of abnormal photoluminescence upon structural phase competence and transition-induced disorder of stable α-FAPbI3. Opt. Mater. Express 2022, 13, 263–271. [Google Scholar]
  22. Monroy, M.I.P.; Goldberg, I.; Elkhouly, K.; Georgitzikis, E.; Clinckemalie, L.; Croes, G.; Annavarapu, N.; Qiu, W.; Debroye, E.; Kuang, Y.; et al. All-evaporated, all-inorganic CsPbI3 perovskite-based devices for broad-band photodetector and solar cell applications. ACS Appl. Electron. Mater. 2021, 3, 3023–3033. [Google Scholar]
  23. Straus, D.B.; Guo, S.; Cava, R.J. Kinetically stable single crystals of perovskite-phase CsPbI3. J. Am. Chem. Soc. 2019, 141, 11435–11439. [Google Scholar] [PubMed]
  24. Chen, B.; Liao, W.; Wu, C. High-stability CsPbIBr2 nanocrystal with nitrogen-doped graphene quantum dot/titanium dioxide for enhancing rhodamine B photocatalytic degradation under visible light. J. Environ. Chem. Eng. 2022, 10, 107534. [Google Scholar]
  25. Far’ain Md Noor, N.; Saiful Badri, M.A.; Salleh, M.M.; Umar, A.A. Synthesis of white fluorescent pyrrolic nitrogen-doped graphene quantum dots. Opt. Mater. 2018, 83, 306–314. [Google Scholar]
  26. Yin, Y.; Cheng, H.; Tian, W.; Wang, M.; Yin, Z.; Jin, S.; Bian, J. Self-assembled δ-CsPbI3 nanowires for stable white light emission. ACS Appl. Nano Mater. 2022, 5, 18879–18884. [Google Scholar]
  27. Steele, J.A.; Jin, H.; Dovgaliuk, I.; Berger, R.F.; Braeckevelt, T.; Yuan, H.; Martin, C.; Solano, E.; Lejaeghere, K.; Rogge, S.M.J.; et al. Thermal unequilibrium of strained black CsPbI3 thin films. Science 2019, 365, 679–684. [Google Scholar]
  28. Kundu, S.; Richtsmeier, D.; Hart, A.; Yeddu, V.; Song, Z.; Niu, G.; Thrithamarassery Gangadharan, D.; Dennis, E.; Tang, J.; Voznyy, O.; et al. Orthorhombic non-perovskite CsPbI3 microwires for stable high-resolution X-ray detectors. Adv. Opt. Mater. 2022, 10, 2200516. [Google Scholar]
  29. Qu, D.; Zheng, M.; Du, P.; Zhou, Y.; Zhang, L.; Li, D.; Tan, H.; Zhao, Z.; Xie, Z.; Sun, Z. Highly luminescent S, N co-doped graphene quantum dots with broad visible absorption bands for visible light photocatalysts. Nanoscale 2013, 5, 12272–12277. [Google Scholar]
  30. Zhang, J.; Zhang, X.; Dong, S.; Zhou, X.; Dong, S. N-doped carbon quantum dots/TiO2 hybrid composites with enhanced visible light driven photocatalytic activity toward dye wastewater degradation and mechanism insight. J. Photochem. Photobiol. A 2016, 325, 104–110. [Google Scholar]
  31. Xu, J.; Wu, Y.; Li, Z.; Liu, X.; Cao, G.; Yao, J. Resistive switching in nonperovskite-phase CsPbI3 film-based memory devices. ACS Appl. Mater. Interfaces 2020, 12, 9409–9420. [Google Scholar] [PubMed]
  32. Yuan, G.; Feng, S.; Yang, Q.; Yi, F.; Li, X.; Yuan, Y.; Wang, C.; Yan, H. Promoting charge separation in a composite of δ-CsPbI3 and covalent organic frameworks. J. Mater. Chem. C 2023, 11, 7570–7574. [Google Scholar]
  33. Sarkar, D.; Ghosh, C.K.; Mukherjee, S.; Chattopadhyay, K.K. Three dimensional Ag2O/TiO2 Type-II (p-n) nanoheterojunctions for superior photocatalytic activity. ACS Appl. Mater. Interfaces 2013, 5, 331–337. [Google Scholar]
  34. Zhang, W.; Xiong, J.; Li, J.; Daoud, W.A. Seed-assisted growth for low-temperature-processed all-inorganic CsPbIBr2 solar cells with efficiency over 10%. Small 2020, 16, 2001535. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, J.; Zhu, H.; Peng, Y.; Li, P.; Chen, S.; Yang, B.; Zhang, J. Photocatalytic performance and degradation pathway of rhodamine B with TS-1/C3N4 composite under visible light. Nanomaterials 2020, 10, 756. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, G.; Xi, Q.; Zhou, H.; Zhao, Y.; Wu, C.; Wang, L.; Guo, P.; Xu, J. Novel inorganic perovskite quantum dots for photocatalysis. Nanoscale 2017, 9, 12032–12038. [Google Scholar]
  37. Qi, X.; Zhang, F.; Chen, Z.; Chen, X.; Jia, M.; Ji, H.; Shi, Z. Hydrothermal synthesis of stable lead-free Cs4MnBi2Cl12 perovskite single crystals for efficient photocatalytic degradation of organic pollutants. J. Mater. Chem. C 2023, 11, 3715–3725. [Google Scholar] [CrossRef]
  38. Cheng, S.; Chen, X.; Wang, M.; Li, G.; Qi, X.; Tian, Y.; Jia, M.; Han, Y.; Wu, D.; Li, X.; et al. In-situ growth of Cs2AgBiBr6 perovskite nanocrystals on Ti3C2Tx MXene nanosheets for enhanced photocatalytic activity. Appl. Surf. Sci. 2023, 621, 156877. [Google Scholar]
  39. Chen, B.; Lu, W.; Xu, P.; Yao, K. Potassium poly(heptazine imide) coupled with Ti3C2 MXene-derived TiO2 as a composite photocatalyst for efficient pollutant degradation. ACS Omega 2023, 8, 11397–11405. [Google Scholar]
  40. Sun, Y.; Long, W.; Guo, Y.; Wei, R.; Wang, Y.; Zhang, J.; Hu, S. Degradation of pollutants by Bi-doped LaFeO3/CQDs/CN Z-scheme heterojunction photocatalysts and mechanism study. Diam. Relat. Mater. 2022, 130, 109555. [Google Scholar]
  41. Preeyanghaa, M.; Vinesh, V.; Neppolian, B. Complete removal of tetracycline by sonophotocatalysis using ultrasound-assisted hierarchical graphitic carbon nitride nanorods with carbon vacancies. Chemosphere 2021, 287, 132379. [Google Scholar] [PubMed]
Figure 1. High-resolution TEM images of (a) NGQDs and (b) δ-CsPbI3. (c) Large-scale and (d) high-resolution TEM images of NGQDs-CsPbI3. (e) Large-scale and (f) high-resolution TEM images of NGQDs-CsPbI3/TiO2. The lattice diagrams are the corresponding SAED patterns.
Figure 1. High-resolution TEM images of (a) NGQDs and (b) δ-CsPbI3. (c) Large-scale and (d) high-resolution TEM images of NGQDs-CsPbI3. (e) Large-scale and (f) high-resolution TEM images of NGQDs-CsPbI3/TiO2. The lattice diagrams are the corresponding SAED patterns.
Molecules 28 07310 g001
Figure 2. (a) SEM image of NGQDs-CsPbI3 microcrystal and EDS elemental mapping of (b) Cs L, (c) Pb L, (d) C K, (e) N K, and (f) I L.
Figure 2. (a) SEM image of NGQDs-CsPbI3 microcrystal and EDS elemental mapping of (b) Cs L, (c) Pb L, (d) C K, (e) N K, and (f) I L.
Molecules 28 07310 g002
Figure 3. (a) XRD patterns of NGQDs, δ-CsPbI3, and NGQDs-CsPbI3. (b) XRD pattern of NGQDs-CsPbI3/TiO2.
Figure 3. (a) XRD patterns of NGQDs, δ-CsPbI3, and NGQDs-CsPbI3. (b) XRD pattern of NGQDs-CsPbI3/TiO2.
Molecules 28 07310 g003
Figure 4. (a) Optical images under sun and UV light of δ-CsPbI3 (left) and NGQDs-CsPbI3 (right) nanocrystals in water. (b) Time-dependent PLs of NGQDs-CsPbI3. (c) The absorption spectra of δ-CsPbI3 and NGQDs-CsPbI3. (d) The PLs of NGQDs-CsPbI3 before and after compositing with TiO2. (e) The PLs of NGQDs-CsPbI3/TiO2 with different NGQD mass ratios. The excitation wavelength is 330 nm.
Figure 4. (a) Optical images under sun and UV light of δ-CsPbI3 (left) and NGQDs-CsPbI3 (right) nanocrystals in water. (b) Time-dependent PLs of NGQDs-CsPbI3. (c) The absorption spectra of δ-CsPbI3 and NGQDs-CsPbI3. (d) The PLs of NGQDs-CsPbI3 before and after compositing with TiO2. (e) The PLs of NGQDs-CsPbI3/TiO2 with different NGQD mass ratios. The excitation wavelength is 330 nm.
Molecules 28 07310 g004
Figure 5. (a) Effects of different samples on the photocatalytic degradation of RhB under visible light (Xe lamp). (b) The corresponding degradation kinetic behaviors. (c) Photocatalytic cycle tests of NGQDs-CsPbI3/TiO2. (d) Effects of scavengers on the catalytic effect of NGQDs-CsPbI3/TiO2.
Figure 5. (a) Effects of different samples on the photocatalytic degradation of RhB under visible light (Xe lamp). (b) The corresponding degradation kinetic behaviors. (c) Photocatalytic cycle tests of NGQDs-CsPbI3/TiO2. (d) Effects of scavengers on the catalytic effect of NGQDs-CsPbI3/TiO2.
Molecules 28 07310 g005
Figure 6. Mechanism of the photocatalytic process of NGQDs-CsPbI3/TiO2, and the inset is the UPS spectra of NGQDs-CsPbI3.
Figure 6. Mechanism of the photocatalytic process of NGQDs-CsPbI3/TiO2, and the inset is the UPS spectra of NGQDs-CsPbI3.
Molecules 28 07310 g006
Table 1. Comparison of dye degradation effects of relevant photocatalysts under visible light.
Table 1. Comparison of dye degradation effects of relevant photocatalysts under visible light.
No.CatalystsDye SolutionEfficiencyRef.
1TiO2RhB in water 10% in 4 h[8]
2CsPbBr3RhB in toluene/ethanol 89% in 100 min[36]
3CsPbCl3RhB in toluene/ethanol 90% in 100 min[36]
4Cs3AgInCl3Sudan Red in ethanol98.5% in 16 min[20]
5Cs4MnBiCl12RhB in ethanol97% in 7 min[37]
6Cs2AgBiBr6/Ti3C2RhB in ethanol100% in 70 min[38]
7NGQDs-CsPbI3/TiO2RhB in water96% in 4 hour
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gu, Y.; Du, X.; Hua, F.; Wen, J.; Li, M.; Tang, T. Nitrogen-Doped Graphene Quantum Dot-Passivated δ-Phase CsPbI3: A Water-Stable Photocatalytic Adjuvant to Degrade Rhodamine B. Molecules 2023, 28, 7310. https://doi.org/10.3390/molecules28217310

AMA Style

Gu Y, Du X, Hua F, Wen J, Li M, Tang T. Nitrogen-Doped Graphene Quantum Dot-Passivated δ-Phase CsPbI3: A Water-Stable Photocatalytic Adjuvant to Degrade Rhodamine B. Molecules. 2023; 28(21):7310. https://doi.org/10.3390/molecules28217310

Chicago/Turabian Style

Gu, Yiting, Xin Du, Feng Hua, Jianfeng Wen, Ming Li, and Tao Tang. 2023. "Nitrogen-Doped Graphene Quantum Dot-Passivated δ-Phase CsPbI3: A Water-Stable Photocatalytic Adjuvant to Degrade Rhodamine B" Molecules 28, no. 21: 7310. https://doi.org/10.3390/molecules28217310

Article Metrics

Back to TopTop