Novel Strategies For The Formulation and Processing of Aluminum Metal Organic Framework Based Sensing Systems Toward Environmental Monitoring of Metal Ions Yongbiao Hua Full Chapter PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

Novel strategies for the formulation and

processing of aluminum metal-organic


framework-based sensing systems
toward environmental monitoring of
metal ions Yongbiao Hua
Visit to download the full and correct content document:
https://ebookmass.com/product/novel-strategies-for-the-formulation-and-processing-o
f-aluminum-metal-organic-framework-based-sensing-systems-toward-environmental-
monitoring-of-metal-ions-yongbiao-hua/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Metal-Organic Frameworks (MOFs) for Environmental


Applications Sujit K. Ghosh

https://ebookmass.com/product/metal-organic-frameworks-mofs-for-
environmental-applications-sujit-k-ghosh/

Rare Earth Metal-Organic Framework Hybrid Materials for


Luminescence Responsive Chemical Sensors Bing Yan

https://ebookmass.com/product/rare-earth-metal-organic-framework-
hybrid-materials-for-luminescence-responsive-chemical-sensors-
bing-yan/

Metal-free synthetic organic dyes Kruger

https://ebookmass.com/product/metal-free-synthetic-organic-dyes-
kruger/

The Rise of Smart Cities : Advanced Structural Sensing


and Monitoring Systems Amir Alavi

https://ebookmass.com/product/the-rise-of-smart-cities-advanced-
structural-sensing-and-monitoring-systems-amir-alavi/
Metal Oxide Powder Technologies: Fundamentals,
Processing Methods and Applications (Metal Oxides) 1st
Edition Yarub Al-Douri (Editor)

https://ebookmass.com/product/metal-oxide-powder-technologies-
fundamentals-processing-methods-and-applications-metal-
oxides-1st-edition-yarub-al-douri-editor/

Metal-Organic Frameworks with Heterogeneous Structures


Ali Morsali

https://ebookmass.com/product/metal-organic-frameworks-with-
heterogeneous-structures-ali-morsali/

Metal Oxides for Optoelectronics and Optics-Based


Medical Applications Suresh Sagadevan

https://ebookmass.com/product/metal-oxides-for-optoelectronics-
and-optics-based-medical-applications-suresh-sagadevan/

Metal oxide-based photocatalysis : fundamentals and


prospects for application Zaleska-Medynska

https://ebookmass.com/product/metal-oxide-based-photocatalysis-
fundamentals-and-prospects-for-application-zaleska-medynska/

Biopolymer-Based Metal Nanoparticle Chemistry for


Sustainable Applications Mahmoud Nasrollahzadeh

https://ebookmass.com/product/biopolymer-based-metal-
nanoparticle-chemistry-for-sustainable-applications-mahmoud-
nasrollahzadeh/
Journal of Hazardous Materials 444 (2023) 130422

Contents lists available at ScienceDirect

Journal of Hazardous Materials


journal homepage: www.elsevier.com/locate/jhazmat

Review

Novel strategies for the formulation and processing of aluminum


metal-organic framework-based sensing systems toward environmental
monitoring of metal ions
Yongbiao hua, Younes Ahmadi, Ki-Hyun Kim *
Department of Civil and Environmental Engineering, Hanyang University, 222 Wangsimni-Ro, Seoul 04763, South Korea

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Application of Al-MOFs are reviewed for


the first-time for metal detection.
• Recent progress in the synthesis/func­
tionalization of Al-MOFs is summarized.
• Various sensing principles of Al-MOFs
sensors are introduced.
• Performance are assessed between Al-
MOFs sensors and other common
sensors.
• The future advancements in this
research field are discussed.

A R T I C L E I N F O A B S T R A C T

Editor: Aluminum is a relatively inexpensive and abundant metal for the mass production of metal-organic frameworks
(MOFs). Aluminum-based MOFs (Al-MOFs) have drawn a good deal of research interest due to their unique
Keywords: properties for diverse applications (e.g., excellent chemical and structural stability). This review has been
Aluminum metal-organic frameworks (Al- organized to highlight the current progress achieved in the synthesis/functionalization of Al-MOF materials with
MOFs) the special emphasis on their sensing application, especially toward metal ion pollutants in the liquid phase. To
Metal ions learn more about the utility of Al-MOF-based sensing systems, their performances have been evaluated for
Sensing

Abbreviations: 3-FSA, 3-formylsalicylic acid ligand. Aptamer 1, 5′ -COOH-(CH2)10-AAA AAA AAA GGG G-SH-3′ ; Aptamer 2, 5′ -TTT TTT AAA ATT TTT T-SH-3′ ;
BET, Brunauer-Emmett-Teller; CAU, Christian-Albrechts-University; CDs, Carbon dots; CNTs, Carbon nanotubes; COF, Covalent-organic framework; CS, Comple­
mentary strand; DUT, Dresden University of Technology; DNA, Deoxyribonucleic acid; DMF, N,N-dimethyl formamide; DPV, Differential pulse voltammetry; ECL,
Electrochemiluminescence; EDS, Energy dispersive spectrometry; GCE, Glassy carbon electrode; H2IPA-V, 5-vinyl isophthalic acid; H2L, 2-(pyrene-1-imine) ter­
ephthalic acid; ICP, Inductively coupled plasma; ICP-MS, Inductively coupled plasma mass spectrometry; ICP-OES, Inductively coupled plasma optical emission
spectrometry; LOD, Limit of detection; MIL, Matérial Institute of Lavoisier; MOF, Metal-organic frameworks; NPs, Nanoparticles; PB, Prussian blue; PEI, Poly­
ethyleneimine; PPy, Polypyrrole; PSM, Post-synthetic modification; QDs, Quantum dots; STY, Space time yield; XRD, X-ray diffraction; ZIF, Zeolite imidazolate
frameworks.
* Corresponding author.
E-mail address: [email protected] (K.-H. Kim).

https://doi.org/10.1016/j.jhazmat.2022.130422
Received 13 September 2022; Received in revised form 11 November 2022; Accepted 15 November 2022
Available online 17 November 2022
0304-3894/© 2022 Elsevier B.V. All rights reserved.
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Performance diverse metallic components in reference to many other types of sensing systems (in terms of the key quality
Post-synthetic modification assurance (QA) criteria such as limit of detection (LOD)). Finally, the challenges and outlook for Al-MOF-based
Nano-particles sensing systems are discussed to help expand their real-world applications.

1. Introduction from low endurance in harsh environments (e.g., high temperature and
moisture) (Ding et al., 2019; Yuan et al., 2018). For example, some
To date, numerous types of sensing systems have been developed for MOFs synthesized by divalent cations (e.g., Zn2+, Cu2+, and Co2+) are
various objectives. For instance, humans can avoid the consumption of often reported to have several disadvantages in terms of chemical
toxins using the olfactory system (which provides the sense of smell), as and/or hydrothermal instabilities (Devic and Serre, 2014).
it allows to differentiate between fresh and expired food (Tricoli et al., To scale up the production of MOFs for diverse applications, it is
2010). Such natural sensors have inspired the scientific community to important to develop new strategies for their fabrication such as the use
design and construct devices that could rapidly determine the quantity of highly abundant, low cost, lightweight, and high valence metals
(and/or quality) of the target components in the surrounding environ­ (Devic and Serre, 2014; Wu et al., 2020). Among a plethora of metal
ment. As a result, various sensing systems (e.g., chemical and biological species (e.g., zirconium, iron, and chromium) present on the earth,
sensors) are now available in diverse fields such as environmental aluminum is one of the most abundant metallic components in the earth
monitoring, food safety, and agriculture applications (Schroeder et al., crust (8.3% by weight) (McLeod and Shaulis, 2018). Moreover, due to
2018; Banica, 2012). the low toxicity and low cost, the selection of highly charged metal
The presence of emerging environmental pollutants (such as metal cationic aluminum (i.e., Al3+) can be favorable in the synthesis of a
ions) poses a significant threat to human health (e.g., endocrine stable and robust MOF through the build-up of strong metal-ligand
disruption, congenital disorder, and carcinogenesis) (Pereira et al., bonds (Devic and Serre, 2014; Wu et al., 2020; Schilling et al., 2016).
2015; Ramírez-Malule et al., 2020; Wilkinson et al., 2017; Lodeiro et al., Further, Al-MOFs can also take the advantage of a green solvent (e.g.,
2019; Ahmed et al., 2021). The application of chemical sensing tech­ water) unlike other MOFs (e.g., zirconium MOFs) generally prepared
nology is considered a promising approach towards the rapid detection with toxic solvents (e.g., DMF) (Schilling et al., 2016; Chen et al.,
of toxic metal ions in bodies of water. Such technology is growing due to 2019a). As such, the ton-scale production of Al-MOFs can be achieved in
its practicality, especially wirh respect to high sensitivity (i.e., detection a water-based route (Gaab et al., 2012). The utility of Al-MOFs is hence
of targets even at low concentrations) and facile operation. Generally, recognized from various fields (such as gas storage, sensing, adsorption,
chemical sensors are composed of sensing and transduction units to separations, and catalysis) with the aid of such meritful properties (Shi
detect and translate the signals into easily readable electrical and/or et al., 2021a; Sikka et al., 2022; Kim et al., 2019; Fan et al., 2019; Wang
optical signals (Paixão and Reddy, 2017). The selection and design of the et al., 2020a; Loiseau et al., 2015) (Fig. 2). Moreover, the good water
sensing materials employed in sensor platforms are thus important to stability and high capacity of Al-MOFs for adsorption in aqueous system
obtain enhanced sensing performance (e.g., sensitivity and stability) was also beneficial toward the sensing application such as metal ions in
(Fang et al., 2018; Paolesse et al., 2017). aqueous system (Samokhvalov, 2018). Al-MOFs can easily be con­
In the construction of an ideal sensing system, the properties of structed with stimuli responsive organic ligands (e.g., 2-aminotereph­
sensing materials are important for obtaining their optimum perfor­ thalic acid) with specific optical characteristics (e.g., luminescence)
mance (e.g., in terms of highly exposed surface, numerous active sites for sensing application (Lin et al., 2020). As such, Al-MOFs chemo­
available for the analytes to bind/react, good mechanical properties, sensors have been developed for the detection of various pollutants (e.g.,
and device flexibility) (Majhi et al., 2022). In this context, metal-organic pharmaceuticals) (Li et al., 2019; Zhang et al., 2019a). For instance, the
frameworks (MOFs) have drawn great attention for sensing applications NH2-MIL-53 (Al) has been utilized for fluorescent detection of phar­
due to their advantageous properties (e.g., large surface areas, cavity maceuticals (e.g., tetracyclines) in liquid milk (Li et al., 2019). The
structures, and tunable porosity) (Fang et al., 2018; Kreno et al., 2012). NH2-MIL-53 (Al) nanosensor exhibited specific recognition toward tet­
MOFs are porous coordination polymers formed by the coordination of racyclines through the formation of hydrogen-bonding interactions be­
metal ions/clusters and organic bridging linkers/ligands (Fig. 1) (Zhou tween the -NH2 of the NH2-MIL-53 (Al) and the –OH of tetracyclines. As
et al., 2012). MOFs with desired properties (e.g., porous nature and
crystalline) can be constructed by adjusting their components (e.g.,
various metal ions and organic ligands) and/or their preparation pro­
cedures (and activation methods) (Lu et al., 2022). Indeed, such flexi­
bilities effectively endow MOFs with great potential advantages for
sensing applications. In addition, the reversibility of MOFs for the
adsorption/desorption of analytes can further ensure their recyclability
(Li et al., 2020a; Liu et al., 2014). The rigid structure of MOFs is helpful
in predicting the formation of interactions between analytes and the
MOF surface at the atomic/molecular level (Li et al., 2020a; Liu et al.,
2014). Despite the many advantages of MOFs, many of them also suffer

Fig. 2. Growth in the annual number of publications on Al-MOFs and their


applications in sensing metal ions(Sources: Web of Science, Google Scholar, and
Scopus databases) based on several key words(e.g., aluminum metal-organic
Fig. 1. Schematic of metal-organic frame works(MOFs). frameworks, metal ions, sensing, and Al-MOFs).

2
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

such, the NH2-MIL-53 (Al) nanosensor was demonstrated to quantify Al-MOF-based sensors for efficient and effective monitoring against a
tetracyclines by inducing the fluorescent quenching (e.g., transferring wide range of environmental pollutants.
the electrons of –NH2/-COOH from NH2-MIL-53 (Al) to the –CO/-OH)
(Li et al., 2019). 2. Formulation of Al-MOF sensors
To date, enormous scientific efforts have been devoted to the syn­
thesis and application (e.g., sorption, catalysis, and membrane) of Al- 2.1. Direct synthesis routes of Al-MOFs
MOFs (Wu et al., 2020; Gaab et al., 2012; Loiseau et al., 2015;
Samokhvalov, 2018; Stock, 2014; Steenhaut et al., 2021; Aguirre-Díaz The structure and properties of Al-MOFs are highly dependent on
et al., 2017). For instance, Samokhvalov introduced the application of their synthesis methods. To date, a variety of direct synthesis strategies
Al-MOFs as advanced sorbent for the removal of various environmental have been reported for the fabrication of Al-MOFs. In the present sec­
pollutants (Samokhvalov, 2018). The utilization of Al-MOFs in mem­ tion, the direct synthesis methods of Al-MOFs (e.g., solvo-/hydrother­
brane techniques has also been dealt in various fields (e.g., gas separa­ mal, reflux method, microwave-assisted, sonochemical,
tion and pervaporation) (Wu et al., 2020). Nonetheless, the application mechanochemical, and electrochemical approach) are discussed in
of Al-MOFs for sensing applications toward metal ions has scantily been relation to their merits and disadvantages (Table 1). In addition, those
reported. As a result, we seek to provide a comprehensive overview on routes for synthesis are also summarized to specify the experimental
sensing application of Al-MOFs against various metal ion pollutants in conditions for diverse case studies (Table 2).
the aqueous system in relation to their synthesized strategies. For this
purpose, we focused on the significance of experimental variables in 2.1.1. Solvo-/hydrothermal method
various methods employed for the synthesis of Al-MOFs (e.g., sol­ The solvothermal method is one of the most commonly used pro­
vo/hydrothermal, sonochemical, and microwave-assisted methods). cedures for the preparation of Al-MOFs (Table 2) (Stock, 2014). This
Further, the merits and disadvantages of such strategies were addressed method employs organic solvents for preparing a suspension of the
for improving their preparation as well as sensing applications. precursors. In case of the hydrothermal method, water is used as a sol­
Up-to-date information on the approaches available for the functional­ vent for suspending the precursors. In this method, the reactions be­
ization of Al-MOFs is also provided to help expand their effective sensing tween organic ligands and inorganic groups of Al-MOFs are usually
applications. The sensing potential of Al-MOFs for metals ions are also carried out in a stainless steel autoclave at autogenous high pressure and
evaluated in terms of quality assurance (e.g., limit of detection (LOD), high temperature (up to 300 ℃) for several hours (or days) (Li et al.,
linear range, and sensing medium). Finally, the current challenge and 2020b). The properties of the Al-MOFs obtained through the sol­
outlook for Al-MOFs in this research field are also outlined. This study is vo-/hydrothermal method significantly depend on the reaction param­
expected to inspire researchers in various fields to develop eters (e.g., pH value, solvent, and reaction temperature). For instance,

Table 1
A summary of Al-MOF synthesis protocols along with their advantages and disadvantages.
No. Synthesis methods Energy sources Reaction Advantages Disadvantages References
time

1 Solvo/hydrothermal Electrical Hours to - Common synthesis of most MOFs, -Requirement of soluble (Steenhaut et al., 2021; Khan et al.,
method heating days - Frequently used for single MOF precursors, 2021; Yang et al., 2016)
crystals, -Long reaction time,
- Easy reaction conditions to control -High energy consumption,
MOF architectures (e.g., size, topology, -High solvent waste
and crystallinity) generation,
-Low reaction rates,
-Not suitable for heat-
sensitive solvents and
reagents
2 Reflux method Electrical Hours to -Simple and less technical instruments, -Long reaction time (Khan et al., 2021; Lenzen et al.,
heating days -Reaction at atmospheric pressure, 2019; Schlu¨sener et al., 2019)
-Low energy consumption,
-Easily synthesis procedure
3 Microwave-assisted Microwave Minutes to -Fast reaction rates, -Difficulty regulating (Vinu et al., 2017; Isaeva et al.,
method irradiation hours -Reduced reaction time, irradiation power, 2015; Khan et al., 2012; Halis et al.,
-Less chemical waste, -Unable to achieve the same 2015)
-Controlled properties of MOFs(in conditions in diverse
terms of porosity and size), instruments,
-Phase selectivity -Concerns about
reproducibility
4 Sonochemical Ultrasonic Minutes to -Fast reaction rates, -Need for probe ultrasonic (Al-Attri et al., 2022; Sud and Kaur,
method wave hours -Simple and energy-saving, instruments, 2021; Kim et al., 2020)
-Ease of operation, -Difficulty in accurately
-Mild reaction condition, controlling reaction
-Easy formation of nanoscale MOFs in temperature
different forms(e.g., sphere
morphology)
5 Electrochemical Electrical Minutes to -Short reaction time, -Undesirable accumulation of (Kalhor et al., 2021; Martinez
method current hours -No side reaction or impurity anions MOF crystals, Joaristi et al., 2012; Ghoorchian
associated with metal salts, -High cost production, et al., 2020)
-Possibility of continuous reactions for -High energy consumption
relatively large scale production,
6 Mechanochemical Mechanical Hours -Simple procedure, -Low crystallinity, (Rubio-Martinez et al., 2017;
method force -Eco-friendly, -Occurrence of side reaction Crawford et al., 2015; Li et al.,
-Organic solvent-free process, 2021)
-Fast reaction rate

3
Y. hua et al.
Table 2
Approaches used for synthesis of Al-MOFs.
No. Al-MOF type Metal sources Organic linkers Reaction Reaction Synthesis Product Remarks Reference
temperature duration medium yield (%)
(℃) (h)

[A] Solvo/
hydrothermal
method
1 CAU-1 AlCl3⋅6 H2O Aminoterephthalic acid 125 5 Methanol – – (Ahnfeldt et al.,
2009)
2 CAU-1-NH2 AlCl3⋅6 H2O Aminoterephthalic acid 115–145 2.6 Methanol – – (Ahnfeldt and
Stock, 2012)
3 CAU-1-(OH)2 AlCl3⋅6 H2O 2,5-dihydroxyterephthalic < 130 4 Methanol – – (Ahnfeldt and
Stock, 2012)
4 CAU-3 Al(NO3)3⋅9 H2O Terterephthalic acid 125 12 Methanol/ – BET: 1550 m2g− 1
(Reinsch et al.,
NaOH 2012a)
5 CAU-3-NH2 Al(NO3)3⋅9 H2O 2-aminoterephthalic acid 125 12 Methanol/ – BET: 1250 m2g− 1
(Reinsch et al.,
NaOH 2012a)
6 CAU-3-NDC AlCl3⋅6 H2O 2,6-naphtalenedicarboxylic acid 130 1 Methanol/ – BET: 2320 m2g− 1
(Reinsch et al.,
NaOH 2012a)
7 CAU-4 Al(NO3)3⋅9 H2O 1,3,5-benzenetrisbenzoate 180 24 DMF 47 BET: 1520 m2g− 1
(Reinsch et al.,
2012a)
8 CAU-6 AlCl3⋅6 H2O 2-aminoterephthalic acid 120 12 2-propanol BET: 620 m2g− 1
(Reinsch et al.,
2012b)
9 CAU-8 Al2(SO4)3⋅18 H2O 4,4′ -benzophenonedicarboxylic 140 12 Water/DMF 40 BET: 600 m2g− 1
(Reinsch et al.,
acid 2013a)
10 CAU-9 Al(NO3)3⋅9 H2O 1,2,4,5-tetrakis-(4- 150 18 Water/DMF – BET: 1118 m2g− 1
(Krüger et al.,
carboxylatophenyl)-benzene 2015)
11 CAU-10 Al(NO3)3⋅9 H2O 1,3-benzene dicarboxylic acid 130 22 Acetonitrile BET: 644 m2g− 1
(Leubner et al.,
4


2020)
12 CAU-10-H Al2(SO4)3⋅18 H2O 1,3-benzene dicarboxylic acid 135 12 Water/DMF – BET: 635 m2g− 1
(Reinsch et al.,
2013b)
13 CAU-10-CH3 Al2(SO4)3⋅18 H2O 5-methylisophthalic acid 130 12 Water/DMF – – (Reinsch et al.,
2013b)
14 CAU-10-OCH3 Al2(SO4)3⋅18 H2O 5-methoxyisophthalic acid 130 12 Water/DMF – – (Reinsch et al.,
2013b)
15 CAU-10-NO2 AlCl3⋅6 H2O 5-nitroisophthalic acid 120 12 Water/DMF – BET: 440 m2g− 1
(Reinsch et al.,
2013b)
16 CAU-10-NH2 AlCl3⋅6 H2O 5-aminoisophthalic acid 120 12 Water/DMF – – (Reinsch et al.,
2013b)
17 CAU-10-OH AlCl3⋅6 H2O 5-hydroxyisophthalic acid 120 12 Water/DMF – – (Reinsch et al.,
2013b)
18 CAU-11 AlCl3⋅6 H2O 4,4 -sulfonyldibenzoic acid 150 12 Water/NaOH 98 BET: 350 m2g− 1
(Reimer et al.,

Journal of Hazardous Materials 444 (2023) 130422


2015)
19 CAU-11-COOH AlCl3⋅6 H2O 3,3′ ,4,4′ - 170 5 Water 50 – (Reimer et al.,
diphenylsulfonetetracarboxylic 2015)
dianhydride
20 CAU-12 Al(NO3)3⋅9 H2O 3,3′ ,4,4′ - 170 5 Water 85 – (Reimer et al.,
diphenylsulfonetetracarboxylic 2015)
dianhydride
21 CAU-13 AlCl3⋅6 H2O Trans-1,4- 130 12 Water/DMF 71 BET: 378 m2g− 1
(Niekiel et al.,
cyclohexanedicarboxylate isomers 2013)
22 CAU-15 Al2(SO4)3⋅18 H2O 1,2-benzenedicarboxylate 120 12 Water/DMF 90 (Reinsch et al.,
2013c)
23 MIL-53 Al(NO3)3⋅9 H2O 1,4-benzenedicarboxylic acid 220 72 Water 41.2 BET: 1203 m2g− 1
(Taheri et al.,
2018)
24 MIL-53 1,4-benzenedicarboxylic acid 200 72 Water 66–84 BET: 710–1064 m2g− 1
(Li et al., 2015)
(continued on next page)
Y. hua et al.
Table 2 (continued )
No. Al-MOF type Metal sources Organic linkers Reaction Reaction Synthesis Product Remarks Reference
temperature duration medium yield (%)
(℃) (h)

Al(OH)3/Al2O3/
boehmite/ Al
(NO3)3⋅9 H2O
25 MIL-53 Al(NO3)3⋅9 H2O 1,4-benzenedicarboxylic acid 220 72 Water – BET: 1140 m2g− 1
(Loiseau et al.,
2004)
2 − 1
26 NH2-MIL-53 AlCl3⋅6 H2O 2-aminoterephthalic 150 24 Water/DMF nearly BET: 1882 m g (Cheng et al.,
acid 100 2013)
27 NH2-MIL-53 Al(NO3)3⋅9 H2O 2-aminoterephthalic acid 130 48–144 DMF – BET: 811 m2g− 1
(Kim et al.,
2012)
28 NH2-MIL-53 AlCl3⋅6 H2O 2-aminoterephthalic acid 130 – Water/DMF 75 (Stavitski et al.,
2011)
29 MIL-69 Al(NO3)3⋅9 H2O 2,6-naphthalenedicarboxylic acid 210 16 Water/KOH 75 – (Loiseau et al.,
2005)
30 MIL-96 Al(NO3)3⋅9 H2O 1,3,5-benzenetricarboxylic acid 210 24 Water – BET: 532 m2g− 1
(Loiseau et al.,
2006)
31 MIL-100 Al(NO3)3⋅9 H2O 1,3,5-benzenetricarboxylic acid 120 12 Water/ethanol – Cetyltrimethylammonium (Seoane et al.,
bromide as surfactant; BET: 2015)
1970 m2g− 1
32 MIL-100 Al(NO3)3⋅9 H2O Trimethyl 1,3,5- 210 72 Water/nitric – BET: 1482 m2g− 1 (Barth et al.,
benzenetricarboxylate acid 2015)
33 MIL-100 Al(NO3)3⋅9 H2O 1,3,5-benzenetricarboxylic acid 200 6 Water/DMF – pH= 1.6–1.7,BET: 1920 m2g− 1
(Yang et al.,
2016)
34 NH2-MIL-101 AlCl3⋅6 H2O 2-aminoterephthalic acid 130 – DMF – – (Stavitski et al.,
2011)
35 NH2-MIL-101 AlCl3⋅6 H2O 2-amino terephthalic acid 130 72 DMF BET: 2100 m2g− 1
(Serra-Crespo
5


et al., 2011)
36 MIL-110 Al(NO3)3⋅9 H2O Trimethyl 1,3,5- 210 72 Water/ nitric – BET: 1408 m2g− 1
(Volkringer
benzenetricarboxylate acid et al., 2007)
37 MIL-116 Al(NO3)3⋅9 H2O Benzene hexacarboxylic 210 24 Water – – (Volkringer
acid et al., 2013)
38 MIL-118 Al(NO3)3⋅9 H2O 1,2,4,5-benzenetetracarboxylic 210 24 Water 66 – (Volkringer
acid et al., 2009a)
39 MIL-120 Al(NO3)3⋅9 H2O 1,2,4,5-benzenetetracarboxylic 210 24 Water/NaOH – pH:12.2, BET: 304 m2g− 1
(Volkringer
acid et al., 2009a)
40 MIL-121 Al(NO3)3⋅9 H2O 1,2,4,5-benzenetetracarboxylic 210 24 Water – pH:1.41, BET: 162 m2g− 1
(Volkringer
acid et al., 2010a)
41 MIL-122 Al(NO3)3⋅9 H2O 1,4,5,8-naphthalenetetracarboxylic 210 24 Water – pH:2.5 (Volkringer
acid et al., 2009b)
42 MIL-129 Al(NO3)3⋅9 H2O 4,40-azobenzenedicarboxylic acid 210 24 Water pH:2.0 (Volkringer

Journal of Hazardous Materials 444 (2023) 130422



et al., 2010b)
43 MOF-253 AlCl3⋅6 H2O 2,2′ -bipyridine-5,5′ -dicarboxylic 120 24 DMF – BET: 2160 m2g− 1
(Bloch et al.,
acid 2010)
44 MOF-467 Al(NO3)3⋅9 H2O 4,4′ ,4′ ’-[benzene-1,3,5-triyl-tris 120 72 HCCOH/DMF 76 BET: 725 m2g− 1
(Wang et al.,
(oxy)]tribenzoic acid 2015a)
45 MOF-519 Al(NO3)3⋅9 H2O 4,4′ ,4′ ’-[benzene-1,3,5-triyl-tris 150 72 DMF/nitric – BET: 2400 m2g− 1
(Gándara et al.,
(oxy)]tribenzoic acid acid 2014)
46 MOF-520 Al(NO3)3⋅9 H2O 4,4′ ,4′ ’-[benzene-1,3,5-triyl-tris 140 96 DMF/formic – – (Lee et al.,
(oxy)]tribenzoic acid acid 2016)
47 DUT-4 Al(NO3)3⋅9 H2O 2,6-naphthalene dicarboxylic acid 120 24 DMF 90.5 BET: 1308 m2g− 1
(Senkovska
et al., 2009)
48 DUT-5 Al(NO3)3⋅9 H2O 4,4′ -biphenyldicarboxylic acid 120 24 DMF 95.6 BET: 1613 m2g− 1
(Senkovska
et al., 2009)
49 [Al3(OH)3(HTCS)2] Al(NO3)3⋅9 H2O 220 72 78 BET: 11 m2g− 1

(continued on next page)


Y. hua et al.
Table 2 (continued )
No. Al-MOF type Metal sources Organic linkers Reaction Reaction Synthesis Product Remarks Reference
temperature duration medium yield (%)
(℃) (h)

tetrakis(4-oxycarbonylphenyl) Water/ (Guo et al.,


silane hydrofluoric 2017)
acid
50 [Al5O2(OH)3(TCS)2(H2O)2] Al(NO3)3⋅9 H2O tetrakis(4-oxycarbonylphenyl) 120 72 Water/formic 63 BET: 1506 m2g− 1
(Guo et al.,
silane acid/DMF 2017)
51 Al(OH)(1,4-NDC)⋅2 H2O Al(NO3)3⋅9 H2O 1,4-naphthalenedicarboxylic acid 180 24 Water 80 pH:2.5 (Comotti et al.,
2008)
52 [Al(OH)(hfipbb)] Al(NO3)3⋅9 H2O 4,4′ -hexafluoroisopropylidene-bis- 160 72 Water/ethanol 94 – (Aguirre-Díaz
(benzoic acid) et al., 2015)
53 [Al(OH)(SDC)] AlCl3 4,4′ -stilbenedicarboxylic acid 180 72 DMF/ acetic BET: 2757 m2g− 1
(Lo et al., 2013)
acid
54 [Al2(OH)2(C16O8H6)] Al(NO3)3⋅9 H2O Biphenyl-3,3′ ,5,5′ -tetracarboxylic 210 72 Water/nitric 75 Piperazine as additive (Yang et al.,
(H2O)6 acid acid 2012a)

[B] Reflux method


55 CAU-10-H Al2(SO4)3/NaAlO2 Sodium isophthalate refluxing temperature 10 Ethanol 95 – (Lenzen et al., 2018
56 CAU-10-H Al2(SO4)3⋅18 H2O 1,3-benzene dicarboxylic acid refluxing temperature 117 Water/DMF 91 BET: 564 m2g− 1 (Fröhlich et al., 2016)
57 CAU-23 NaAlO2⋅0.2H2O 2,5-thiophenedicarboxylic acid refluxing temperature 6 Water 54 BET: 1056 m2g− 1 (Schlüsener et al., 2020)
58 CAU-23 AlCl3/NaAlO2 2,5-thiophenedicarboxylic acid refluxing temperature 6 Water/NaOH 84 – (Lenzen et al., 2019)
59 MIL-68 AlCl3⋅6 H2O 1,4-benzene dicarboxylic acid 125 18 DMF – BET: 976 m2g− 1
(Tehrani and Zare-Dorabei, 2016b)
60 MIL-68 AlCl3⋅6 H2O 1,4-benzene dicarboxylic acid 130 18.5 DMF – – (Yang et al., 2012b)
61 MIL-68 AlCl3⋅6 H2O 1,4-benzene dicarboxylic acid 125 18 DMF – BET: 976 m2g− 1 (Tehrani and Zare-Dorabei, 2016a)
62 NH2-MIL-68 AlCl3⋅6 H2O 2-aminoterephthalic acid 130 18 DMF – BET: 1170.9 m2g− 1
(Rahmani et al., 2022)
63 MIL-160 AlCl3⋅6 H2O 2,5-furandicarboxylic acid 100 24 Water/NaOH – BET: 1170 m2g− 1 (Cadiau et al., 2015)
64 MIL-160 AlCl3 2,5-furandicarboxylic acid 100 24 Water/NaOH 55 BET: 1178 m2g− 1 (Tannert et al., 2019)
6

65 MIL-160 AlCl3⋅6 H2O 2,5-furandicarboxylic acid 100 24 Water/NaOH 57 – (Henry and Samokhvalov, 2022)
66 MIL-160 AlCl3 2,5-furandicarboxylic acid 100 24 Water 93 BET: 1150. m2g− 1 (Permyakova et al., 2017)
67 DUT-5 AlCl3 20-sulfone-biphenyl-4,40-dicarboxylic acid 178 72 Diethylformamide – BET: 1260 m2g− 1 (Raja et al., 2015)
68 [Al(OH)(C6H3NO4)]⋅nH2O Al2(SO4)3⋅18 H2O 2,5-pyrpyrroledicarboxylic acid 120 12 Water/NaOH 93 BET: 1130 m2g− 1 (Cho et al., 2020)

[C] Microwave-
assisted method
69 CAU-13 AlCl3⋅6 H2O Trans-1,4-cyclohexanedicarboxylic acid 130 0.75 Water/DMF – – (Niekiel et al., 2014)
70 MIL-53 Al 1,4-benzenedicarboxylic acid 200 0.77 Water 69 MW power/frequency: 1200 W/ (Vinu et al., 2017)
(NO3)3⋅9 H2O 2.45 GHz, BET: 1403 m2g− 1
71 MIL-53 Al 1,4-benzenedicarboxylic acid 130 3 Water/DMF 68–86 MW power: 200–400 W, BET: (Sun et al., 2022)
(NO3)3⋅9 H2O 716–990 m2g− 1
72 NH2-MIL-53 Al 2-aminoterephthalic acid 200 0.77 Water 73 MW power/frequency: 1200 W/ (Vinu et al., 2017)
(NO3)3⋅9 H2O 2.45 GHz, BET: 865 m2g− 1
73 MIL-96 Al 1,3,5-benzenetricarboxylic acid 210 1 Water MW power: 400 W, BET: 216 m2g− 1
(Khan et al., 2012)

Journal of Hazardous Materials 444 (2023) 130422



(NO3)3⋅9 H2O
74 MIL-100 Al Trimesic acid trimethyl ester 200 1.38 Water/nitric 67 MW power/frequency: 1200 W/ (Vinu et al., 2017)
(NO3)3⋅9 H2O acid 2.45 GHz, BET: 1701 m2g− 1
75 MIL-100 Al Trimethyl 1,3,5-benzenetricarboxylate 210 0.02–0.08 Water/nitric – MW power: 400 W, BET: 1056 m2g− 1
(Khan et al., 2012)
(NO3)3⋅9 H2O acid
76 MIL-100 Al Trimesic acid 210 0.08 Water – MW power: 1200 W, BET: 1050 m2g− 1
(Moreau et al., 2020)
(NO3)3⋅9 H2O
2 − 1
77 MIL-100 Al Trimesic acid 210 0.5 Water/nitric 98 MW power: 1200 W, BET: 2400 m g , (García Márquez et al.,
(NO3)3⋅9 H2O acid pH: 0.7 2012)
78 NH2-MIL-101 AlCl3⋅6 H2O 2-aminoterephthalic acid 130 6 DMF – MW power: 300 W, BET: 1980 m2g− 1, (Haque et al., 2014)
79 NH2-MIL-101 AlCl3⋅6 H2O 2-aminoterephthalic acid 130 0.17–0.5 DMF – MW power/frequency: 12 W/6 GHz, (Isaeva et al., 2015)
BET: 1980 m2g− 1,
80 MIL-110 Al Trimethyl 1,3,5-benzenetricarboxylate 210 0.03 Water/ – MW power: 1200 W, BET: 639 m2g− 1, (Khan et al., 2012)
(NO3)3⋅9 H2O NaOH
(continued on next page)
Y. hua et al.
Table 2 (continued )
81 DUT-4 Al 2,6-napthalenedicarboxylic acid 120 0.77 Ethanol 91 MW power/frequency: 1200 W/ (Vinu et al., 2017)
(NO3)3⋅9 H2O 2.45 GHz, BET: 1184 m2g− 1
82 DUT-4-(NO2)2 AlCl3⋅6 H2O 1,5-dinitro-3,7-naphthalenedicarboxylic 140 4 DMF – BET: 578 m2g− 1 (Halis et al., 2015)
acid
83 DUT-5 Al 4,4′ -biphenyldicarboxylic acid 120 0.77 Ethanol 76 MW power/frequency: 1200 W/ (Vinu et al., 2017)
(NO3)3⋅9 H2O 2.45 GHz, BET: 1157 m2g− 1
84 DUT-5-NO2 AlCl3⋅6 H2O 3-nitro-4,4′ -biphenyldicarboxylic acid 150 3 DMF – BET: 1677 m2g− 1 (Halis et al., 2015)
85 DUT-5-NH2 AlCl3⋅6 H2O 3-amino-4,40-biphenyldicarboxylic acid 150 3 DMF – BET: 1966 m2g− 1 (Halis et al., 2015)
86 DUT-5-SO2 Al 5,5-dioxo-dibenzo[b,d]thiophen-3,7- 160 1 DMF – BET: 1530 m2g− 1 (Halis et al., 2015)
(ClO4)3⋅9 H2O dicarboxylic acid
87 [Al2(OH)2(C16O8H6)] AlCl3⋅6 H2O Biphenyl-3,3′ ,5,5′ -tetracarboxylic acid 210 0.17 Water 83.5 MW power: 300 W, BET: 1272 m2g− 1
(Thomas-Hillman et al.,
(H2O)6 2019)

[D] Sonochemical method


88 MIL-53 Al(NO3)3⋅9 H2O 1,4-benzenedicarboxylic acid 150 6 Water – Sonication power: 24 kHz for 1 h, BET: 872 m2g− 1 (Al-Attri et al., 2022)
89 MIL-53 Al(NO3)3⋅9 H2O 1,4-benzenedicarboxylic acid 220 24 Water – Sonication power: 24 kHz for 0.5 h, BET: 1538.6 m2g− 1 (Ahadi et al., 2022)
7

90 Al fumarate Al2(SO4)3⋅18 H2O Fumaric acid – 0.02–0.5 Water/NaOH – Sonication power: 500 W and 20 kHz, BET: 1100–1169 m2g− 1
(Kim et al., 2020)

[E] Electrochemical
method
91 MIL-53 Aluminum Terephthalic acid 20–90 – Water/DMF – KCl as electrolyte, current density: 100 mA cm− 2, BET: (Martinez Joaristi et al.,
plates 1243 m2g− 1 2012)
92 NH2-MIL- Aluminum 2-aminoterephthalic acid room 0.5 Water/ – KNO3 as electrolyte, current density: 10 mA cm− 2, BET: (Kalhor et al., 2021)
53 plates temperature ethanol 205 m2g− 1
93 NH`-MIL- Aluminum 2-aminoterephthalic acid 90 – Water/DMF – KCl as electrolyte, current density: 10 mA cm− 2, BET: (Martinez Joaristi et al.,
53 plates 788 m2g− 1 2012)
94 MIL-100 Aluminum 1,3,5-benzenetricarboxylic > 60 1 Water/ – KCl as electrolyte, current density: 10 mA cm− 2 (Martinez Joaristi et al.,
plates acid ethanol 2012)

[F] Mechanochemical method


95 Al(fumarate)OH) Al2(SO4)3⋅18 H2O Fumaric acid 150 – – – NaOH as co-reactant, screw speed: 55 rpm, feed rate: 10 g min− 1, BET: 1010 m2g− 1
(Crawford et al., 2015)

Journal of Hazardous Materials 444 (2023) 130422


Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

controlling the pH in a hydrothermal reaction significantly alters the process (e.g., crystals growth) for synthesis of MOF materials.
topologies of the resulting Al-MOFs (Volkringer et al., 2010a). In such
reactions, aluminum sources (e.g., Al(NO3)3⋅9 H2O), organic ligands (e. 2.1.2. Reflux method
g., pyromellitic acid (C6H2(CO2H)4), and solvent (i.e., deionized water) As an eco-friendly and facile approach, the reflux method has also
are combined to react in an autoclave under various pH conditions to been used for the fabrication of Al-MOFs (Khan et al., 2021; Cadiau
obtain different topologies of Al-MOFs (Volkringer et al., 2010a). Under et al., 2015; Henry and Samokhvalov, 2022; Lenzen et al., 2019; Tehrani
acidic (low) pH, the formation of MIL-118 and MIL-121 occurred and Zare-Dorabei, 2016a). The synthesis of MOFs based on the reflux
favorably due to the connectivity of aluminum cations in the inorganic method involves a variety of solvents (e.g., water, ethanol, and N,
subnetwork via μ2-hydroxo corners of AlO6 octahedra (Volkringer et al., N-dimethylformamide (DMF)) to induce chemical reactions in a reflux
2010a). At higher pH, the aluminum-centered octahedra were linked to condenser (at a temperature of 100–130 ℃) (Table 2) (Henry and
each other through common edges of two μ2-hydroxo groups to form Samokhvalov, 2022; Lenzen et al., 2018; Tannert et al., 2019). Various
MIL-120. In addition, the number of coordinating carboxylic oxygen types of Al-MOF materials such as CAU-10-H, CAU-23, MIL-160, and
atoms per pyromellitate ligands increased from 4 to 8 by adjusting the MIL-68 have been synthesized using the reflux method (Schlüsener
pH value (from 1.4 to 12.2), which provided more sites to connect et al., 2020; Tehrani and Zare-Dorabei, 2016b; Fröhlich et al., 2016). For
aluminum ions (Volkringer et al., 2010a) (Fig. 3). instance, the compound [Al(C6H2O4S)(OH)]⋅xH2O (named CAU-23) was
The selection of proper solvent and reaction temperature is also vital formed by a chemical reaction between AlCl3 and NaAlO2 as the
for the production of desired Al-MOF products. Such a relationship can aluminum sources and sodium thiophenedicarboxylate (Na2TDC) as the
be observed from the production of NH2-MIL-53 Al-MOFs using the organic linkers in a reflux method (Lenzen et al., 2019). The as-prepared
solvothermal method (i.e., aluminum chloride as a metal salt and amino- CAU-23 exhibited nano-scale particle rods (200 nm) with a BET surface
terephthalic acid as an organic ligand) (Stavitski et al., 2011) (Fig. 4). area of 1250 m2g− 1 (Fig. 5a and b). The rod shape was formed by
The presence of organic solvent (e.g., dimethylformamide) promoted alternating units of four consecutive trans and cis corner-sharing AlO6
quick dissolution of organic linker (i.e., amino-terephthalic acid) to polyhedra. Four trans and cis corner-sharing AlO6 polyhedra corre­
yield a metastable product (i.e., NH2-MOF-235(Al)) at comparatively sponded to the straight and helical sections of the Al-O chain, respec­
low temperatures (Stavitski et al., 2011). In contrast, the slow dissolu­ tively. (Fig. 5(c-e)) (Lenzen et al., 2019). Likewise, MIL-160(Al)
tion rate of organic linker (amino-terephthalic acid) resulted in the materials (surface area of 1150 m2g− 1) were prepared on a large-scale
formation of highly stable products NH2-MIL-53 at intermediate tem­ of 400 g through the reflux method (Permyakova et al., 2017). This
peratures when the synthesis was performed using water as solvent. approach nonetheless seemed to require long reaction times (hours to
Temperature also exerts a noticeable effect on the morphology of Al- days). Despite some flaws, the reflux method can be regarded as a po­
MOFs (e.g., MIL-53(Al)) (Taheri et al., 2018; Al-Attri et al., 2022). For tential technique for the large-scale synthesis of MOFs with many ad­
instance, Al(NO3)3⋅9 H2O and 1,4-benzenedicarboxylic acid were mixed vantageous properties (e.g., simple technical instruments, reaction at
with a molar ratio of 5:2.5 in 25 ml of water. The hydrothermal reaction atmospheric pressure, low energy consumption, and convenient syn­
was conducted for 72 h at various temperatures (e.g., 150 and 220 ℃). thesis procedure) (Khan et al., 2021).
Particle sizes of MIL-53(Al) were observed in the range of 0.77–2.19 µm
at 150 ℃. In contrast, they were reduced to 0.68–1.36 µm at 220 ℃ to 2.1.3. Microwave-assisted method
reflect the effects of increasing of solubility, nucleation, and crystal The microwave-assisted method relies on the use of electromagnetic
growth at higher temperature (Al-Attri et al., 2022). Despite various radiation (frequency range: 300 MHz – 300 GHz) to directly heat the
advantageous aspects of the solvo(hydro)thermal method (e.g., pro­ reaction medium (Sud and Kaur, 2021). It is considered a time and en­
duction of good quality MOF crystals and prevention of the volatilized ergy efficient approach for the fabrication of Al-MOFs (Stock, 2014;
toxic substance during reaction), the major drawback of the solvo Stock and Biswas, 2012; Vinu et al., 2017). As compared to the external
(hydro)thermal method is the energy and time (hours to days) required. heating method (e.g., conventional electrical heating), the
Solvo(hydro)thermal techniques also suffer from unobservable reaction microwave-assisted method is an efficient option to rapidly and

Fig. 3. Influence of pH on the formation of various Al-MOF materials(Volkringer et al., 2010a), copyright 2010, American Chemical Society (ACS) publishing group.

8
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 4. Formation mechanism of NH2-MOF-235(Al), NH2-MIL-101(Al), and NH2-MIL-53(Al) in DMF/water mixtures using solvo(hydro)thermal method (Stavitski
et al., 2011), copyright 2011, Wiley-VCH publishing group.

uniformly increase the temperature throughout the reaction medium aluminum-benzenetricarboxylates (Al-BTCs) was regulated by varying
(Sud and Kaur, 2021). As such, the microwave-assisted method was the reaction time of microwave assisted synthesis approach (Khan et al.,
reported to efficiently increase the rate of the nucleation/growth process 2012). As shown in the XRD pattern and the product weight graph of
for Al-MOFs (Stock and Biswas, 2012). Therefore, the time for the in­ MIL-96 (Fig. 6d and e), highly porous MOFs (e.g., MIL-100) were ac­
duction period and complete crystallization of microwave-assisted quired at short reaction times (e.g., 1 min). In contrast, increases in the
CAU-1-(OH)2 was shortened considerably (e.g., 10–180 min) relative reaction time (>5 min) caused the gradual conversion of MIL-100 (less
to conventional solvothermal method (e.g., 10–280 min) depending on stable) into MIL-96 (more stable) (Khan et al., 2012). The short reaction
the reaction temperature (Fig. 6a) (Ahnfeldt et al., 2011). In another time might favor the formation of kinetically controlled MOFs, while the
report, the microwave method was also reported to shorten the synthesis long period promotes the production of thermodynamically stable
time of NH2-MIL-101(Al) significantly from tens of hours (conventional MOFs.
solvothermal method) to 10–30 min (Isaeva et al., 2015). Microwave Microwave power is another critical parameter to consider in the
irradiation is effective for the reduction of synthesis time while serving determination of the morphological features of Al-MOF materials. The
as an effective route for determining the crystal size. As exemplified by effect of microwave power on the production of MIL-53(Al) was inves­
the work of Isaeva’s group, the crystal sizes of Al-MOFs obtained tigated (Sun et al., 2022). In this study, Al(NO3)3⋅9 H2O, 1,4-benzenedi­
through a microwave-assisted approach was smaller (e.g., 50–100 nm in carboxylate acid, and DMF/H2O were used as the aluminum source,
case of NH2-MIL-101(Al)) compared to the solvothermal method (from 2 organic ligand, and co-solvent, respectively. The mean crystal size of the
to 3 µm) (Isaeva et al., 2015). The small crystal size was mainly due to obtained MIL-53(Al) was reduced from 13.26 to 10.19 nm when the
fast nuclei crystallization and growth during microwave reactions applifed microwave power increased from 200 W to 400 W (Fig. 6g-h).
(Fig. 6b and c) (Isaeva et al., 2015). Accordingly, the control of key Correspondingly, the BET surface area and pore volume decreased from
parameters (such as reaction time and microwave power) should 990 to 716 m2g− 1 and 0.77–0.63 cm3g− 1, respectively. Such observa­
contribute to the precise growth of crystalline Al-MOFs with tions were accounted for by high microwave power (e.g., in terms of the
well-defined geometries. aggravation of particles agglomeration, influence of nucleation, and
The regulation of the reaction time in the microwave-assisted crystal rate for MIL-53(Al) crystals) (Sun et al., 2022). Despite the ad­
method is also an effective option to obtain Al-MOF materials with vantages of microwave-assisted methods, the rising concerns of repro­
desired properties (e.g., stable and high porosity or different pore sizes) ducibility could still be the limiting factor for its wide application (Sud
(Khan et al., 2012). For instance, the porosity of and Kaur, 2021; Zhang et al., 2022a).

9
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 5. The preparation of CAU-23 materials using reflux method: (a) SEM image, (b) nitrogen sorption measurement, (c) reconstructed 3D reciprocal lattice, (d) the
repetition of cis and trans corner-sharing AlO6 polyhedra forming the inorganic building unit of CAU-23, and (e) the full structure of CAU-23 (Lenzen et al., 2019),
copyright 2019, Springer nature publishing group.

2.1.4. Sonochemical method method was also found to be higher than those of the conventional
The sonochemical method has been demonstrated as an elegant hydrothermal approach (BET surface: 1078 m2 g− 1) (Kim et al., 2020).
synthesis route compared to the common routes (e.g., solvo/hydro­ Such a sonochemical method was also employed for the fabrication of
thermal method) in terms of simplicity, short reaction time, ease of other Al-MOFs (e.g., MIL-53) (Ahadi et al., 2022). Using ultrasonication,
operation, mild reaction condition (e.g., ambient temperature), and the synthesized MIL-53 also showed a higher surface area (1538.6 m2
energy-efficiency (Stock and Biswas, 2012; Safaei et al., 2019; Lee et al., g− 1) than that of the hydrothermal method (1140 m2 g− 1) (Ahadi et al.,
2013). The sonochemical method involves the production of bubbles in 2022; Loiseau et al., 2004). Despite the advantageous aspects of the
the reaction medium under high energy ultrasonic radiation sonochemical method (e.g., shorten reaction time and mild reaction
(20 kHz − 10 kHz) (Sud and Kaur, 2021; Stock and Biswas, 2012). The condition), only a few studies are exclusively concerned with the
process governing the generation, growth, and collapse of bubbles is sonochemical synthesis of Al-MOFs (Kim et al., 2020). More efforts
referred to as cavitation, which leads to the rapid release of energy with should be made to expand the utilization of the sonochemical method
a high heating/cooling rates (1010 Ks− 1), extreme temperature (5000 K), for the preparation of Al-MOF materials.
and high pressure (1000 atm) inside the cavitation zone as well as in its
vicinity (around 200 nm) (Li et al., 2020b; Stock and Biswas, 2012). This 2.1.5. Electrochemical method
whole process is useful for promoting chemical reactions and immediate The electrochemical method for the preparation of MOFs is based on
formation of nuclei for crystallization at ambient temperature (Fig. 7a) applying an electric current between two electrodes immersed in an
(Kim et al., 2020). For this purpose, a solution containing an aluminum electrolyte (Fig. 8a) (Lee et al., 2013). The metal ions are continuously
source and organic ligands was treated using an ultrasonic probe to yield supplied through the metal anodic dissolution of the electrode by elec­
highly crystallized Al-based MOFs (Fig. 7b). As displayed in Fig. 7c-d, trochemical oxidation. Subsequently, MOFs are formed by the reactions
the Al fumarate MOF prepared by the sonochemical method exhibited between these metal ions and the dissolved molecules of organic linker
good sphere-like morphology. In contrast, an Al fumarate MOF synthe­ present in the electrolyte medium. Besides, protic solvents are used to
sized by a conventional hydrothermal method showed continuously avoid metal deposition on the electrodes (i.e., cathode). As this method
inter-grown agglomerates of spherical particles. Such good sphere-like can technically avoid the inclusion of anions (e.g., nitrate and chloride)
morphology of Al fumarate MOF was attributed to the prevention of from metal salts, it is advantageous for the large-scale production of pure
particle agglomeration by ultrasonication. As such, the BET surface area crystalline Al-MOFs. Furthermore, the method also benefits from several
(1169 m2 g− 1) of Al fumarate MOF prepared by the sonochemical advantages such as short reaction time (e.g., within minutes or hours),

10
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 6. The preparation of Al-MOF materials using microwave-assisted method: (a) comparison of induction and reaction time between microwave (grey) and
solvothermal (black) method for production of CAU-1-(OH)2 (Ahnfeldt et al., 2011), copyright 2011, Wiley-VCH publishing group. SEM images of the NH2-MIL-101
(Al) synthesized by various methods: (b) solvothermal method and (c) microwave-assisted method (Isaeva et al., 2015), copyright 2015, Springer nature publishing
group. The influence of reaction time on the production of Al-BTCs: (d) XRD patterns, and (e) obtained weight of Al-BTCs (Khan et al., 2012), copyright 2012,
Elsevier publishing group. The effect of microwave power on the synthesis of MIL-53(Al): (f) 200 W, (g) 300 W, and (h) 400 W (Sun et al., 2022), copyright 2022,
Elsevier publishing group.

mild reaction conditions (at ambient temperature and pressure), and area of MIL-53(Al)-NH2 was further improved to 788 m2 g− 1 using a
scalability (Raptopoulou, 2021; Rubio-Martinez et al., 2017; Zhang high reaction temperature (90 ℃), high solubility solvent for linkers
et al., 2019b). (DMF), and deprotonation reagent of linkers (tributylmethylammonium
In a typical electrochemical synthesis procedure, aluminum metallic methyl sulfate) (Martinez Joaristi et al., 2012). In addition, this elec­
plates were inserted in the presence of organic linkers (e.g., 2-amino trochemical technique can help formulate Al-MOFs with uniform
terephthalic acid) and an electrolyte support (e.g., potassium nitrate) structures (e.g., 3D cauliflower morphology) (Fig. 8c) (Kalhor et al.,
in an ethanol/water (10/90, v/v) mixture under stirring (Kalhor et al., 2021). Of late, limited literature has reported on the electrochemical
2021). MIL-53(Al)-NH2 was successfully synthesized (Fig. 8b) on the synthesis of aluminum-based MOFs (Kalhor et al., 2021; Martinez
basis of the electrochemical method. Upon establishing a current density Joaristi et al., 2012).
of 10 mA cm− 2, the hydroxide ions were in-situ electro-generated on the
surface of the cathode to increase the pH at the cathode surface (Fig. 8b). 2.1.6. Mechanochemical method
As such, water molecules were electro-reduced to deprotonate ligands. The mechanochemical method is an eco-friendly strategy for the
In addition, the anode surface was oxidized to generate aluminum cat­ large-scale production of Al-MOF materials. This approach is mainly
ions required for the formation of MIL-53(Al)-NH2 crystals. In the applied to utilize MOF precursors with poor solubility (e.g., oxides,
electrochemical approach, pore-blocking of prepared Al-MOF was pre­ hydroxides, and carbonates) (Rubio-Martinez et al., 2017; Chen et al.,
vented due to the absence of cationic salts (required to construct MOF) 2019b). The mechanochemical method relies on both physical and
(Kalhor et al., 2021). The electrochemically formulated Al-MOFs (e.g., chemical interactions using external mechanical force/energy (e.g.,
MIL-53(Al)-NH2) generally exhibit high surface area (e.g., 204.75 m2 milling, grinding, and extrusion) (Stock and Biswas, 2012;
g− 1) with a wide pore size (of 4.2 nm) (Kalhor et al., 2021). The surface Rubio-Martinez et al., 2017; Crawford and Casaban, 2016). Under such

11
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 7. The preparation of Al-MOFs materials using sonochemical method: (a) schematic illustration of preparing aluminum fumarate MOF materials, (b) PXRD
patterns of aluminum fumarate MOF materials prepared by various methods: sonochemical method (black) and conventional hydrothermal method (red). SEM
images of aluminum fumarate MOF materials prepared by various methods: (c) sonochemical method and (d) conventional hydrothermal method (Kim et al., 2020),
copyright 2020, Elsevier publishing group.

high mechanical energy and dry conditions (i.e., without solvents), the 2.2. Functionalization of Al-MOFs with sensing elements
intramolecular bonds in metal salts are mechanically disrupted to
reconstruct new chemical bonds between metal cations and organic li­ A number of strategies (e.g., utilization of stimuli responsive organic
gands (Younis et al., 2021). For example, the mechanochemical method ligands during the direct synthesis of Al-MOFs, post-synthetic modifi­
has been used to synthesize Al(fumarate)(OH) MOF materials by cation (PSM) approaches, and formation of Al-MOFs composites) can be
external extrusion (Fig. 9a) (Crawford et al., 2015). In such cases, re­ considered to impart or improve the sensing properties of Al-MOF ma­
actants (i.e., cation source, organic ligands, and anion (OH-) source) terials, as discussed below.
were mixed in the absence of solvent (Fig. 9b) (Crawford et al., 2015).
The product obtained through mechanochemical method possessed BET 2.2.1. Utilization of functional organic ligands in direct synthesis
surface areas (1010 m2 g− 1) comparable to those of other methods (e.g., The basic information on the construction and luminescent mecha­
solvothermal method, 1020 m2 g− 1) (Gaab et al., 2012; Crawford et al., nisms of luminescent MOFs has been explained in elsewhere (Lustig
2015; Jeremias et al., 2014). Due to the lack of solvents, rapid reactions, et al., 2017; Hu et al., 2014; Yang et al., 2021). The luminescent prop­
and small free volume present in the twin screw extruder, the mecha­ erties of MOFs can be accounted for by the combined effects of several
nochemical method showed an excellent process efficiency in the syn­ factors such as organic ligands-metal charge transfers, the organic li­
thesis of Al(fumarate)(OH) MOF materials (e.g., high space time yield gands with aromatic moieties or extended π systems, and introduced
(STY: 27000 kg m− 3 day− 1) relative to a solvothermal-based process lanthanide ions (Cui et al., 2012). Luminescent Al-MOF-based sensors
(STY: 3600 kg m− 3 day− 1)) (Crawford et al., 2015). Unlike traditional have commonly been built with the aid of stimuli responsive organic
methods (e.g., solvothermal approach), the main benefits of the mech­ ligands to effectively gain the capacity for the luminescent detection
anochemical synthesis include its environmentally friendly nature, high towards analytes (e.g., metal ions) (Li et al., 2019; Qin et al., 2021; Yang
production yield of Al-MOF in a short reaction period, lack of organic et al., 2013). In this respect, the luminescent properties of Al-MOFs can
solvent, and cost-effectiveness (Rubio-Martinez et al., 2017). However, be mainly attributed to the emission phenomena of organic ligands and
it should be noted that this method may also suffer from the generation ligand-metal charge transfer (Lin et al., 2020; Yang et al., 2013; Liu and
of some by-products due to the interference of anionic salts. Hence, Yan, 2016). Some examples of π-containing organic ligands are 2-amino­
further purification is frequently required to improve the quality of the terephthalic acid, terephthalic acid, 2,2′ ,2′ -(-triazine-2,4,6-triimino)tri­
final products (Rubio-Martinez et al., 2017). benzoic acid), 5-formyl isophthalic acid, and 2,5-thiophenedicarboxilic
acid (Li et al., 2019; Zhu et al., 2020; Díaz-Ramírez et al., 2020; Ma et al.,
2021; Nandi et al., 2019). The luminescence properties of Al-MOFs
greatly rely on the π electrons of organic ligands, which help generate

12
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 8. The preparation of Al-MOF materials using electrochemical method: (a) typical schematic of preparing MOFs (Lee et al., 2013), copyright 2013, Springer
nature publishing group. (b) Fabrication of MIL-53(Al)-NH2 and (c) their SEM images (Kalhor et al., 2021), copyright 2021, Springer nature publishing group.

excitation of π electrons (π→π * ) of ligands (Lin et al., 2020). It is also


suggested that the luminescence in Al-MOFs generates from the charge
transfer between metal ions and organic ligands (Zhang et al., 2019a;
Yang et al., 2013). It is recognized that the ligand-metal charge transfer
should be based on the electronic transition (e.g., from an organic
linker-localized orbital to a metal-centered orbital) (Cui et al., 2012). As
a good example, the blue emission of NH2-MIL-53 (Al) at 450 nm under
the excitation of 340 nm was attributed to the electronic transition from
the 2-aminoterephthalic acid ligands to Al metal ions clusters (Xie et al.,
2021). Further, ratiometric fluorescence can decrease the environmental
interference and vision tiredness to provide easy-to-differentiate color
and intensity changes compared with that of single-color fluorescence
detection (Chen et al., 2020). As such, the construction of ratiometric
fluorescent Al-MOF-based sensors can be realized through the combi­
nation of stimuli responsive and optically active ligands. In this regard, a
ratiometric fluorescent ruthenium-based Al-MOF (Ru@MIL− NH2)
sensor was developed using stimuli responsive 2-aminoterephthalic acid
organic ligands and Ru(bpy)3Cl2⋅6 H2O] as optically active organic li­
gands (Yin et al., 2017). Despite the facile introduction of sensing ele­
ments into a MOF architecture through a direct method, these sensors
generally suffer from challenges including low linker solubility, poor
thermal stability, and undesirable chemical reactions between cations
and linker functional moieties (Schneemann et al., 2014; Sun and Zhou,
2015; Cohen, 2012).
Fig. 9. The preparation of Al(fumarate)(OH) MOF materials using mechano­
chemical method: (a) chemical reaction and (b) the twin screw extruder in­
2.2.2. Post-synthetic modification
strument used (Crawford et al., 2015), copyright 2015, Royal Society of
The post-synthetic modification (PSM) approaches are promising
Chemistry (RSC) publishing group.
options to functionalize Al-MOFs with sensing moieties. PSM helps
introduce new functional groups within the main molecular skeleton of
the signals for the detection process (Lin et al., 2020; Hu et al., 2014; Qin
synthesized Al-MOFs while maintaining their intrinsic structure (Cohen,
et al., 2021). For example, MIL-101(Al)-NH2 MOFs were prepared using
2012; Mandal et al., 2021; Kamel et al., 2021; Zhang and Yan, 2019;
2-aminoterephthalic acid as an organic ligand (Lin et al., 2020). The
Wang and Cohen, 2009; Yin et al., 2019; Kalaj and Cohen, 2020). Based
formulated Al-MOFs exhibited luminescence properties with the

13
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

on the nature of bonds formed between modified reagents and MOFs, the exchange between the MOFs and the external metal ions (cation ex­
PSM strategy can be divided into two types, i.e., dative and covalent change), or a combined mechanism (e.g., cations exchanges and
PSMs (Cohen, 2012). ligand-analytes interactions)) (Dao et al., 2018; Liu et al., 2019; Huangfu
The dative PSM method is generally used to introduce luminescent et al., 2021; Wang et al., 2017). The quenching effect can be quantita­
compounds (e.g., lanthanide ions such as samarium (Sm3+), europium tively obtained using the Stern-Volmer equation (Huangfu et al., 2021;
(Eu3+), dysprosium (Dy3+), erbium (Er3+), and ytterbium (Yb3+)) to the Johnson et al., 1998):
structure of Al-MOFs for sensing applications (Younis et al., 2021; Cui
I0/I=1+Ksv[M] (1)
et al., 2014). In fact, lanthanide ions can coordinatively react with the
uncoordinated active sites of the organic linkers in MOFs to form where I0 and I are the luminescent intensities of MOFs before and
functional luminescent Al-MOFs materials (Zhang and Yan, 2019; Yan, after recognition of the analyte, respectively, [M] is the concentration of
2017). The lanthanide ions can exhibit (characteristic) narrow 4–4 f the analyte, and Ksv is the Stern-Volmer constant. A higher Ksv value
transitions emission (Cui et al., 2012). As such, the luminescent prop­ indicates a higher efficiency of the MOF sensor (Huangfu et al., 2021). In
erties of lanthanide/Al-MOFs materials should be generated from the f-f the case of ligand-analyte interactions, the exterior metal ions/analytes
transitions of lanthanide ions (Luo et al., 2019; Dao et al., 2018). More coordinated with the organic ligands can change the organic properties
specifically, the Eu3+@MOF-253 can exhibit several strong emission (e.g., electron density, rigidity, or conformation), thereby resulting in
peaks locating at 579, 592, 612, and 699 nm, which correspond to the luminescent quenching of MOFs (Huangfu et al., 2021; Hao et al., 2013).
5
D0→7FJ (J=0–4) transition of Eu3+ ions (Luo et al., 2019). For example, a turn-off luminescent sensor with pyrene functionalized
On the other hand, covalent PSM of Al-MOFs is achieved through the ligand (MIL-53-L) was designed and constructed for sensing Cu2+ in an
formation of covalent bonds between functional reagents (e.g., 3-formyl­ aqueous solution (Liu and Yan, 2016). The imine part of the organic
salicylic acid ligand (3-FSA)) and the organic linkers of Al-MOFs (Kamel ligand (i.e., 2-(pyrene-1-imine) terephthalic acid (H2L)) offered binding
et al., 2021). For example, a luminescent Al-MOF (e.g., 3-FSA-NH2-­ sites for Cu2+ ions, while its pyrene moieties acted as fluorophores. The
MIL-53(Al)) was fabricated for the detection of metal ions via a Schiff Cu2+ ions effectively quenched the luminescence emission at 567 nm of
base reaction. The strong covalent bond between the sensing element MIL-53-L relative to other metal ions (e.g., Hg2+, Zn2+, Mn2+, and Fe2+)
and the Al-MOF offered high chemical stability (robust framework) to under excitation at 337 nm. In such cases, N atoms of the internal imine
the resulting materials (Kamel et al., 2021). groups interacted with Cu2+ ions to convert them into amide groups.
MIL-53-L showed a linear quenching range of 0–500 µM with a high Ksv
2.2.3. Formation of Al-MOF composites value of 6.15 × 103 µM− 1 towards the detection of Cu2+ ions (Liu and
Composite formulations are alternative functionalization methods to Yan, 2016).
endow Al-MOFs with strong and specific sensing abilities (e.g., electro- The cation exchange mechanism in luminescent MOFs having a
chemiluminescence and electrochemical) (Feng et al., 2021, 2020; neutral framework is based on the exchange of the analyte metal ions
Wang et al., 2020b; Duan et al., 2019). Quantum dots (QDs) as novel and post-modified framework of MOFs (Yang et al., 2013; Huangfu
semiconductor materials are effective options to offer high quantum et al., 2021; Dao and Ni, 2017). For instance, the cation exchange be­
yield and good electrical conductivity (Aguilera-Sigalat and Bradshaw, tween Fe3+ and the framework Al3+ ions in MIL-53(Al) caused fluores­
2016). In addition, Prussian blue (PB) consisting of Fe4[Fe(CN)6]3- ex­ cence quenching at 410 nm emission in MIL-53(Al) under an excitation
hibits unique catalytic properties (Duan et al., 2019). Carbon nanotubes wavelength at 305 nm (Fig. 10a) (Yang et al., 2013). Such a phenome­
(CNTs) have the unique properties of excellent conductivity (Camilli and non was attributed to the transformation of strong-fluorescent MIL-53
Passacantando, 2018). (Al) to weak-fluorescent MIL-53(Fe) (Fig. 10b). As a result, this MIL-53
Al-MOF-based composite sensors are generally synthesized through (Al) sensor allowed sensitive detection of Fe3+ ions in an aqueous so­
the incorporation of nanomaterials/nanofillers (e.g., CdS quantum dots, lution with a linear range of 3–200 µM (LOD of 0.9 µM) (Yang et al.,
Au nanoparticles (NPs), carbon dots (CDs), CdTe quantum dots, poly­ 2013). In another study, post-synthetically modified Al-MOFs with Tb3+
pyrrole, carbon nanotubes (CNTs), and Prussian blue (PB)) on the sur­ ions (i.e., MIL-110(Al)/Tb3+) were constructed to sense Fe3+ ions (Dao
face or into the cavities of MOFs (Feng et al., 2021, 2020; Wang et al., and Ni, 2017). The fluorescence response of MIL-110(Al)/Tb3+ toward
2020b; Duan et al., 2019; Xu and Yan, 2016). For example, target metal ions was validated for its potential use for metal ion
electro-chemiluminescent Al-MOF-based composite sensors were detection. Fig. 10c shows that only Fe3+ ions can fully quench the
formulated through the incorporation of CdS quantum dots (QDs) and fluorescence emission in MIL-110(Al)/Tb3+, implying its high Fe3+ ion
Au nanoparticles as electro-chemiluminescence emitters (energy donor) selectivity. Herein, MIL-110(Al)/Tb3+ was used for detection of Fe3+
and energy acceptors, respectively (Feng et al., 2021). In another study, ions with a wide linear range (0–500 µM), low detection limit (655 nM),
carbon nanotubes and Prussian blue were combined with MIL-53(Al), and high quenching constant (Ksv: 6.177 ×103 µM− 1) using 301 nm
and the MIL-53(Al)/CNTs/PB composite was able to generate a cur­ excitation (Fig. 10d). As confirmed by EDS and ICP technologies, the
rent response because of the electrochemical catalytic activity of PB luminescent quenching of MIL-110(Al)/Tb3+ toward Fe3+ ions was
(Duan et al., 2019). accounted for by cation exchange between Fe3+ and Tb3+ ions. In
addition, the MIL-110(Al)/Tb3+ exhibited high selectivity against Fe3+
3. Applications and sensing mechanism of Al-MOF sensors for ions in aqueous solution due to the stronger affinity of the –COOH group
detection of metal ions to Fe3+ ions (Dao and Ni, 2017).
In some cases, the fluorescence quenching phenomena are complex
In general, Al-MOF-based sensors are built to detect the target ana­ and may involve the combination of several quenching mechanisms
lytes (e.g., metal ions) based on luminescent, electro­ (Lustig et al., 2017; Gao et al., 2018; Das et al., 2018). For example,
chemiluminescence, and electrochemical techniques. The strategies for aqueous phase sensing of bismuth ions (Bi3+) using [Al8(O­
using functionalized Al-MOFs for sensing of various metal ions are H)4(OCH3)8(BDC(OH)2)6]⋅x H2O (CAU-1-(OH)2) was achieved based on
summarized in Table 3. cation exchange and the ligand-analyte interaction mechanisms (Gao
The luminescent-based Al-MOF sensors exhibited variable lumines­ et al., 2018). CAU-1-(OH)2 emits bright green fluorescence at 505 nm
cent emission intensity, i.e., luminescence enhancement (turn-on) and upon excitation of 365 nm in water due to the intramolecular proton
quenching (turn-off) when recognizing the analytes (Shi et al., 2021b). transfer process between the ligand and water molecule (Jayaramulu
Generally, the quenching luminescence mechanism of MOFs by metal et al., 2010). Further, among various metal ions (e.g., Na+, K+, Ba2+, and
ions includes (i) the interactions between the organic ligands and the Ca2+), the specificity of CAU-1-(OH)2 towards Bi3+ ions was explained
exterior metal ions/analytes (ligand-analytes interaction) and (ii) cation

14
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Table 3
Summary of Al-MOF-based sensors employed for the detection of various metal ions.
No. Al-MOF materials Sensing approach/type Sensing Sensing Dynamic LOD Selectivity* Reference
mechanism media range
μM μM
[A] Fe3þ ions
1 MIL-53(Al) Luminescence/turn off Cation exchange Aqueous 3 − 200 0.9 K+, Na+, Ca2+, Mg2+, Cu2+, (Yang
solution Zn2+, Co2+, Mn2+, Ni2+,Hg2+, et al.,
Cd2+, Pb2+, Fe2+, Cr3+, and 2013)
Al3+
2 Al-MIL-53-N3 Luminescence/turn off Cation exchange/ Water – 0.03 Na+, K+, Ag+, Ca2+, Pb2+, (Das et al.,
ligand-analyte Mn2+,Fe2+, Co2+, Ni2+, Cu2+, 2018)
interactions Zn2+, Cd2+, Hg2+, Al3+, and
Cr3+
3 MIL-53(Al)/Eu3þ Luminescence/turn off Cation exchange Aqueous – 0.793 K+, Na+,Sr2+, Fe2+, Mg2+, (Dao et al.,
solution Hg2+, Zn2+, Pb2+, Cd2+, Ni2+, 2018)
Co2+, Mn2+, Cu2+,Fe3+, In3+,
and Cr3+
4 MIL-110(Al)/ Luminescence/turn off Cation exchange Aqueous 0–500 0.655 K+, Na+, Mg2+, Pb2+, Sr2+, (Dao and
Tb3þ solution Zn2+, Cd2+, Hg2+, Sn2+, Ni, 2017)
Mn2+, Ni2+, Co2+, Fe2+, Cu2+,
Cr3+, and In3+
5 Eu3þ@MIL- Luminescence/turn off Cation exchange Aqueous 0.5–500 0.5 K+, Na+, Ag+, Ca2+,Mg2+, (Zhou
53–COOH(Al) solution Cu2+, Zn2+, Co2+, Ni2+, Hg2+, et al.,
Cd2+, Pb2+, Fe2+, and Cr3+ 2014)
6 Al-MOF Luminescence/turn on Electron transfer Water 1–7 6.62 Li+, K+, Cu2+, Zn2+, Ca2+, (Ma et al.,
(tricarboxylate Mn2+, Cd2+, Pb2+, Hg2+, 2021)
ligands) Cr3+, Sb3+, Na+, Cu+, Fe2+,
Mg2+, Ba2+, Co2+, Ni2+, Sn2+,
Fe3+, and Sr2+
[B] Cu2þ ions
7 MIL-53-L Luminescence/turn off Ligand-analyte Aqueous 0–500 – Hg2+, Zn2+, Mn2+, Fe2+, (Liu and
interactions solution Co2+, and Pb2+ Yan, 2016)

8 Eu @MOF-253 Luminescence/turn off Ligand-analyte Aqueous 0–100 0.66 Na+, Mg2+, Al3+, K+, Ca2+, (Luo et al.,
interactions solution Cr3+, Mn2+, Co2+, Ni2+, Zn2+, 2019)
Cd2+, and Pb2+
9 NH2-MIL-101(Al) Luminescence/turn off Ligand-analytes Aqueous 1.5–625 0.17 Li+, Na+, K+, Ca2+, Mg2+, (Zhang
@ZIF-8 interactions solution Mn2+, Co2+, Ni2+, Hg2+, et al.,
Pb2+, and Cd2+ 2018b)
3
10 NH2-MIL-53(Al)/ Electrochemical Differential pulse Aqueous 0.015–6.294 3.84 × 10− Cd2+, Hg2+, Ca2+, Ni2+, and (Wang
Ppy voltammetry solution Zn2+ et al.,
2020b)
[C] Hg2þ ions
3
11 NH2-MIL-53(Al) Luminescence/turn off – Aqueous 0.002–38.27 0.23 × 10− Zn2+,Ca2+, Mn2+, Cd2+, (Ren et al.,
solution Mg2+, Na+, K+, Ni2+, Pb2+, 2020)
Cr3+, Fe2+, Fe3+, Co2+, rutin,
Al3+, sucrose, glucose, and
palmatine
12 NH2-MIL-53(Al) Luminescence/turn off Ligand-analyte Aqueous 1–17.3 0.15 Li+, K+, Na+, Ni2+, Co2+, (Zhang
interactions solution Ca2+, Pb2+, Cd2+,Zn2+, Mn2+, et al.,
and Mg2+ 2019a)
5 1 6
13 MIL-53(Al) Electrochemiluminescence – Aqueous 10− - 10− 4.1 × 10− Na+, Ca2+, K+, Ag+, Co2+, (Feng et al.,
@CdTe-PEI solution Al3+, Mg2+, Cu2+, Fe3+, and 2020)
Zn2+
14 Eu3þ/ Ratiometric and colorimetric Coordination Water 0.065–150 13 × 10− 3
Zn2+, Ni2+, Na+, Mn2+, Mg2+, (Xu and
CDs@MOF-253 luminescence/turn off interactions K+, Pb2+, Cu2+, Fe3+,Ca2+, Yan, 2016)
Cd2+, Fe2+, Al3+, Cr3+, and
Co2+
[D] Pb2þ ions
4 5
15 MIL-53(Al) Electrochemiluminescence – Aqueous 10− - 10 2.4 × 10− Na+, Ca2+, K+, Ag+, Co2+, (Feng et al.,
@CdTe-PEI solution Al3+, Mg2+, Cu2+, Fe3+, and 2020)
Zn2+
3
16 NH2-MIL-53(Al)/ Electrochemical Differential pulse Aqueous 0.005–1.93 1.52 × 10− Cd2+, Hg2+, Ca2+, Ni2+, and (Wang
Ppy voltammetry solution Zn2+ et al.,
2020b)
[E] Pd2þ ions
19 CAU-10-V-H Luminescence/turn off Ligand-analyte Water – 0.11 Ag+, Cr3+, Cu2+, Cd2+, Co2+, (Nandi
interactions Eu3+, Fe2+, Mg2+, Pb2+, et al.,
Mn2+, Ni2+, Ru3+, and Zn2+ 2020)
20 Al-CAU-10- Luminescence/turn off Ligand-analyte Water – 0.148 Ag+, Cr3+, Cu2+, Cd2+, Co2+, (Ghosh
OC3H5 interactions Eu3+, Fe2+, Mg2+, Pb2+, et al.,
Mn2+, Ni2+, Al3+, Zn2+, Na+, 2021)
K+, Pb2+, and Hg2+
[F] Agþ ions
17 Sm3þ@MIL-121 Luminescence/turn on Antenna effect Aqueous 0–500 0.09 Na+, K+, Mg2+,Ca2+, Al3+, (Hao and
solution Mn2+, Fe3+, Co2+, Ni2+, Cu2+, Yan, 2015)
Ag+, Zn2+, Cd2+, and Pb2+
(continued on next page)

15
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Table 3 (continued )
No. Al-MOF materials Sensing approach/type Sensing Sensing Dynamic LOD Selectivity* Reference
mechanism media range
μM μM

18 Eu @MIL-121 Luminescence/turn on Antenna effect Aqueous 0–100 0.1 Na+, K+,Mg2+, Ca2+, Al3+, (Hao and
solution Cr3+, Mn2+, Fe3+, Co2+, Ni2+, Yan, 2014)
Cu2+, Ag+, Zn2+,Cd2+, Hg2+,
and Pb2+
[G] Bi3þ ions
21 CAU-1-(OH)2 Luminescence/turn off Cations exchange/ Aqueous 0–100 2.16 Na+, K+, Ba2+, Ca2+, Cu2+, (Gao et al.,
ligand-analyte solution Co2+, Cd2+,Ni2+, Zn2+, Mn2+, 2018)
interactions Cr3+, Al3+, and Bi3+
[H] Fe2þ ions
22 MOF-253 Luminescence/turn off Ligand-analyte Aqueous 5–100 0.5 Na+, Ba2+, Ca2+, Mg2+, Pd2+, (Lu et al.,
interactions solution Zn2+,Cd2+, Ni2+, Co2+, Cu2+, 2014)
Cr3+, and Fe3+

*The selectivity is the ability of sensors to recognize the target analytes in the presence of interferences.

Fig. 10. The application of MIL-53-L for Cu2+ sensing: (a) schematic of MIL-53-L for sensing Cu2+ ions, (b) the relative intensities of MIL-53-L in different metal ions
aqueous solution and (c) the possible conformational changes of MIL-53-L for detecting Cu2+ ions (Liu and Yan, 2016), copyright 2016, Elsevier publishing group.

by the intramolecular proton transfer mechanism. Such a mechanism is absorbed energy to the luminescent probe (e.g., luminescent MOFs) as
ascribed to the coordination of Bi3+ ions and hydroxyl groups of an effective option to enhance the emission intensity of luminescent
CAU-1-(OH)2 through which fluorescence quenching is induced (Gao probes (Lustig et al., 2017). As an example, a lanthanide ion (Ln3+)@Al
et al., 2018). It was also found that the fluorescence emission in (OH) (H2btec)⋅H2O (Ln3+@MIL-121) fluorescent probe was designed
CAU-1-(OH)2 was quenched by Bi3+ ions in water due to the following through the post-modification method (for Ag+ detection) (Hao and
possible reasons: (i) parts of Al3+ in the inorganic clusters replaced by Yan, 2015). After post-synthetic functionalization of Ln3+ (e.g., Sm3+),
Bi3+ caused a change in –COO-Al3+ to –COO-Bi3+ and (ii) Bi3+ ions Ag+/Sm3+@MIL-121 showed intense characteristic peaks of Sm3+ due
strongly interacted with the hydroxyl groups of CAU-1-(OH)2, thereby to the antenna effect of Ag+ ions. A typical energy transfer of
weakening the intramolecular proton transfer process (Fig. 11a). The Ag+/Sm3+@MIL-121 was found as follows. First, the lanthanide-organic
sensitivity assay demonstrated a linear relationship between the fluo­ ligand complex was excited from the singlet S0 ground state to the
rescence emission (505 nm) in CAU-1-(OH)2 and the concentration of singlet S1 excited state upon absorbing the energy. Then, the energy of
(0–1 ×10− 4 M) Bi3+ ions under 365 nm excitation (Fig. 11b). The cor­ the S1 excited state was transferred to the triplet T1 excited state of the
responding LOD for the fluorescence quenching assay was determined to ligands. Further, the energy of triplet T1 excited state was transferred to
be 2.16 µM, highlighting its great potential for real-world applications the excited f states of the Ln3+ ions (e.g., Sm3+). Finally, the luminescent
(Gao et al., 2018). emission (f-f) of lanthanide Sm3+ ions was generated (Hao and Yan,
The luminescent emission in MOFs can be further enhanced by the 2015; Feng and Zhang, 2013)(Fig. 12a). During the above energy pro­
sensitization process (also called the antenna effect) (Yan, 2017). Spe­ cess, the antenna effect of Ag+ ions functioned as follows: (i) the Ag+
cifically, analytes acting as strong photon absorbers transferred the ions reduced the rate constant for non-radiative deactivation and energy

16
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 11. The application of Al-MOF for Fe3+ ion sensing based on cation exchange. The application of MIL-53(Al) for detection of Fe3+ ions: (a) schematic of the
cation exchange mechanism for MIL-53(Al) to sense Fe3+, (b) fluorescence spectra of MIL-53(Al) and MIL-53(Fe) (Yang et al., 2013), copyright 2013, ACS publishing
group. The application of MIL-110(Al)/Tb3+ for detection of Fe3+ ions: (c) luminescent intensity of MIL-110(Al)/Tb3+ against various metal ions in the aqueous
solution. (d) The relative intensity of MIL-110(Al)/Tb3+ as a function of Fe3+ ions concentration (Dao and Ni, 2017), copyright 2017, RSC publishing group.

loss of lanthanide complexes during the S0→S1 energy process. This was
confirmed by Ag+/Sm3+@MIL-121 having a longer lifetime (97.6 µs)
than those of Sm3+@MIL-121 (18.0 µs). (ii) The perturbation of the
spin-orbit coupling in Ag+ ions (heavy-atom effect) promoted inter­
system crossing of ligand S1→T1 energy transfer (Liu et al., 2012; Pellow
and Vala, 1989). (iii) The outer-shell d electrons of Ag+ caused
metal-to-ligand charge transfer, which resulted in the sensitization of
fluorescent Sm3+. As such, Sm3+@MIL-121 exhibited excellent selec­
tivity for Ag+ as compared to other metal ions (e.g., Na+, K+, and Pb2+)
(Fig. 12b). All three processes described above helped Ag+ ions
contribute to the effective energy transfer from ligands to lanthanide
ions, thereby leading to the fluorescent enhancement of
Ag+/Sm3+@MIL-121. Furthermore, the Sm3+@MIL-121 exhibited a fast
detection time (< 1 min) for the detection of Ag+ ions since the
Ag+-induced fluorescence reaction was a rapid process (Fig. 12c). The
fluorescence intensity at 603 nm emission of Sm3+@MIL-121 also
showed an increase in the linear relationship with increasing Ag+ con­
centrations (0–500 µM) under excitation at 320 nm (Fig. 12d) (Hao and
Yan, 2015).
Recently, electrochemiluminescence (ECL) was found to be a
promising method for metal ion sensing due to its advantages including
low background and rapid response (Feng et al., 2020). ECL involves the
generation of electroactive species through electrochemical reaction at
Fig. 12. The application of CAU-1-(OH)2 for Bi3+ ion sensing: (a) the mecha­ an electrode surface. Such active species undergo high-energy electro-­
nism of CAU-1-(OH)2 for the detection of Bi3+ (green: Al; orange: Bi; gray: C; transfer reactions (i.e., reaching excited state) to emit light (Hua et al.,
white: H; red: O) and (b) linear fitting results of fluorescence versus the con­ 2022). In this regard, sensitive ECL aptasensors for multiple determi­
centration of Bi3+ ions (Gao et al., 2018), copyright 2018, Elsevier publish­ nation of Hg2+ and Pb2+ were designed using MIL-53(Al)@CdTe-PEI
ing group. and Au NP/Pt NP labelled aptamers as ECL signal probes and target
recognition probes, respectively (Fig. 13a) (Feng et al., 2020). The
occurrence of enhanced resonance energy transfer was confirmed by the

17
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 13. The application of Sm3+@MIL-121 for sensing Ag+ ions: (a) schematic of the fluorescent enhancement of Sm3+@MIL-121 by Ag+ ions. (b) Relative
fluorescence intensities of Sm3+@MIL-121 against various metal ions. (c) Variation of fluorescent intensity of Sm3+@MIL-121 at various immersion times. (d)
Fluorescence intensity of Sm3+@MIL-121 as a function of Ag+ ion concentration (Hao and Yan, 2015), copyright 2015, RSC publishing group.

overlap in the emission spectra of the CdTe quantum dots (QDs) and 53(Al)@CdTe-PEI: 5 µL, scan rate: 100 mV s− 1, aptamer 1-Pt NPs: 14 µL,
absorption spectra of Au or Pt NPs (Fig. 13b). In addition, the MIL-53(Al) aptamer 2-Au NPs: 8 µL) (Fig. 14c) (Feng et al., 2020). More recently, a
and CdTe significantly enhanced the current flow in the sensor for Cu2+/Pb2+ ions electrochemical sensor composed of NH2-MIL-53
sensing metal ions due to their high electrical conductivity. As for the (Al)/polypyrrole (PPy) composites was designed (Wang et al., 2020b).
Hg2+ ions detection (path 1), the Hg2+ ions were initially recognized by NH2-MIL-53(Al) and polypyrrole (PPy) composites were formed through
T-rich oligonucleotides (aptamer 2-Au NPs) (Fig. 13a). Then, aptamer π-π interactions between the PPy and NH2-MIL-53(Al). Such strong in­
2-Au NPs fell off the ECL electrode after the formation of hairpin DNA teractions gave the NH2-MIL-53(Al)/PPy sensor good stability (90% of
structures (T-Hg2+-T base pair). As a result, the dual surface plasmon initial response remained after 2 weeks). As a result of the synergistic
resonance effect of Au and Pt NPs was replaced with the single surface effects between NH2-MIL-53(Al) and PPy, the NH2-MIL-53(Al)/PPy
plasmon effect of Pt NPs to reduce the ECL signal (Eq. 5). The detection sensor exhibited suitable sensing performance toward Cu2+ and Pb2+.
of Pb2+ ions was accompanied by the formation of Pb2+-G-quadruplexes Furthermore, the NH2-MIL-53(Al)/PPy sensor recorded high selectivity
(path 2 and 3). As such, the Pt NPs were effectively stacked on the towards Cu2+ and Pb2+ ions in the presence of interferent ions (e.g.,
electrode, thereby quenching the ECL intensity of the ECL aptasensor Cd2+, Ca2+, and Zn2+), which can be attributed to the high affinity of
through enhanced resonance energy transfer between Pt NPs and CdTe. amino and carboxyl groups of NH2-MIL-53(Al) for Cu2+/Pb2+ ions
In other words, the original energy needed for the excitation of (Wang et al., 2020b). Despite the good performance (e.g., low detection
CdTe* was dissipated by Pt NPs as CdTe* emitted fewer photons. The limit) of electrochemiluminescence or electrochemical based Al-MOFs,
corresponding ECL processes can be listed as follows: only a few studies have been carried out to assess their sensing capac­
ity toward metal ions as listed in Table 3 (Feng et al., 2020; Wang et al.,
CdTe + e- → CdTe-• (2) 2020b).
S2O2-
8 +e →-
SO2-
4 + SO-•
4 (3)
4. Performance evaluation of Al-MOF-based sensing systems
CdTe +-•
SO-•
4 → CdTe* + SO2-
4 (4)
4.1. Overall performance evaluation of Al-MOF-based sensors
CdTe* → CdTe + hv (5)

MIL-53(Al)@CdTe-PEI based ECL sensor was used to detect metal The sensing performances (e.g., in terms of LOD, dynamic detection
ions with a dynamic concentration range under optimal conditions (MIL- range, and selectivity) of various Al-MOF based sensors against various

18
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Fig. 14. The application of MIL-53(Al)@CdTe-PEI based ECL sensor for Pb2+/Hg2+ ion detection: (a) schematic of MIL-53(Al)@CdTe-PEI based ECL sensor for
Pb2+/Hg2+ ions detection (1: path 1, 2: path 2, and 3: path 3). (b) UV–vis spectra of Pt NPs, Au NPs, and CdTe, (c) ECL emission measured at different concentrations
of Hg2+ (path 1) and Pb2+ (path 2 and 3) (Feng et al., 2020), copyright 2020, Elsevier publishing group.

metal ions are summarized in Table 3. In this section, we compare the background noise (Li et al., 2017). Interestingly, luminescent-based
performance of different types of Al-MOF-based sensors against various Al-MOFs were widely reported for the detection of metal ions over
metal ions (e.g., Cu2+, Hg2+, and Pb2+ ions) mainly in terms of LOD as a other types of sensors (e.g., electrochemical and electro­
key quality assurance metric. In this regard, the top five performers have chemiluminescence) (Table 3). The good universality of
been selected in terms of LOD, regardless of the metal species and/or luminescent-based Al-MOFs might be attributed to their simple formu­
sensing principle. Then, the best sensing systems are also selected and lations, ease of operation, and cost-effectiveness.
discussed for each metal species based on the same criteria (i.e., LOD). The best performers in terms of lowest LOD value for each metal
Accordingly, the top five performers for the detection of metal ions, if species were selected for evaluation. In the case of Fe3+ ions, the Al-MIL-
organized in terms of LOD value, were: MIL-53(Al)@CdTe-PEI (Hg2+, 53-N3-based luminescent sensor was found to have the best LOD
4.1 ×10− 6 µM) > MIL-53(Al)@CdTe-PEI (Pb2+, 2.4 ×10− 5 µM) > NH2- (0.03 µM) (Das et al., 2018). Its superior performance may be attributed
MIL-53(Al) (Hg2+, 0.23 ×10− 3 µM) > NH2-MIL-53(Al)/Ppy (Cu2+, to two factors including: (i) the interaction of electron-withdrawing –N3
3.84 ×10− 3 µM) (Feng et al., 2020; Wang et al., 2020b; Ren et al., 2020). groups of 2-azido-1,4-benzenedicarboxylic acid ligands with
Here, the highest sensing ability of the MIL-53(Al)@CdTe-PEI-based electron-deficient Fe3+ ions and (ii) partial cation exchange between
electrochemiluminescent sensor against Hg2+ was attributed to the Al3+ and Fe3+ ions during the sensing process. The synergic effect of
sensitive and specific recognition element (aptamer 2) toward Hg2+ both processes endows the Al-MIL-53-N3-based luminescent sensor with
ions. Besides, the utilization of CdTe quantum dots endowed the Hg2+ the highly sensitive detection of Fe3+ ions in aqueous media (Das et al.,
ion sensor with strong ECL emission due to its good electrical conduc­ 2018). In addition, the aluminum CAU-1-(OH)2-based luminescent
tivity and high quantum yield. The high specific surface area of MIL-53 sensor showed the best Bi3+ ion sensing performance with a LOD value
(Al) acted as good supporter for CdTe quantum dots (Feng et al., 2020). (2.16 µM) (Gao et al., 2018). Here, the carboxylate and hydroxyl groups
Besides, such MIL-53(Al)@CdTe-PEI-based electrochemiluminescent in CAU-1-(OH)2 can strongly coordinate with Bi3+ ions. The high spe­
sensors also provided the best sensing performance (2.4 ×10− 5 µM) cific surface area (1386 m2 g− 1) of CAU-1-(OH)2 offered a large number
toward Pb2+ ions in an aqueous solution when using aptamer 1 as a of active sites for sensing Bi3+ ions (Gao et al., 2018). Likewise, because
recognition element (Feng et al., 2020). The sensing performance of of the large pore properties of MOF-253, Fe2+ ions can easily enter into
Al-MOF-based electrochemiluminescent sensors was generally notice­ the voids to react with the bipyridine group of MOF-253 (Lu et al.,
able in the detection of divalent metal ions (e.g., Pb2+ and Hg2+) over 2014). As such, the aluminum MOF-253-based luminescent sensor
other Al-MOF-based sensors (e.g., electrochemical and luminescent offered the best sensitivity (LOD: 0.5 µM) for detecting Fe2+ ions (Lu
sensors) (Table 3). Such results might reflect the highly sensitive char­ et al., 2014).
acteristics of electrochemiluminescent techniques using the low The best Cu2+ ion sensing system was identified as the NH2-MIL-53

19
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

(Al)/Ppy-based electrochemical sensor (LOD: 3.84 ×10− 3 µM) (Wang Ag+ ions (Hao and Yan, 2015). Such an antenna effect enhanced the
et al., 2020b). The performance of this Cu2+ ion sensing platform re­ energy transfer from ligands to Sm3+. As such, the
flected the synergy between high specific area of the NH2-MIL-53(Al) Sm3+@MIL-121-based luminescent sensor can be highly sensitive (LOD:
and the excellent conductivity of the PPy nanowires that boosted active 0.09 µM) for detecting Ag+ ions (Hao and Yan, 2015). In the case of Pd2+
sites for the detection of Cu2+ ions (Wang et al., 2020b). Further, the ion detection, the CAU-10-V-H-based luminescent sensor showcased the
nitrogen on the pyrrole unit of PPy nanowires can effectively coordinate lowest LOD value (0.11 µM) among all Pd2+ sensing systems (Nandi
with Cu2+ ions. The amino and carboxylic groups from NH2-MIL-53(Al) et al., 2020). Such superior performance of the CAU-10-V-H-based
had metal ions chelation properties for Cu2+ ions. As such, NH2-MIL-53 luminescent sensor should be ascribable to the strong interactions be­
(Al)/Ppy-based electrochemical sensors can efficiently detect Cu2+ ions tween Pd2+ ions and vinyl functional groups of 5-vinyl isophthalic acid
at the lowest LOD level of 3.84 × 10− 3 µM based on the differential pulse (H2IPA-V) from CAU-10-V-H (Nandi et al., 2020). The cyclic utilization
voltammetry (DPV) technique (Wang et al., 2020b). In addition, the best of this sensor was also examined up to five cycles after thorough washing
Ag+ sensing system with a low LOD value of 0.09 µM was achieved using in each experimental cycle. Accordingly, about 70% of reduction in the
a Sm3+@MIL-121-based luminescent sensor due to the antenna effect of sensing quenching efficiency occurred after five cycles due to the

Table 4
Comparison between Al-MOF-based sensor and non-Al-MOF-based sensor for sensing various metal ions.
No. Sensors Sensing media LOD Linear range Reference

µM µM

[A] Fe3þ ions


1 Al-MIL-53-N3-based luminescent sensor Water 0.03 – (Das et al., 2018)
2 Eu-MOF-based luminescent sensor Aqueous solution 0.39 0.01–220 (Guo et al., 2020)
3 [In(L)(m2-OH)]0.5 H2O-based luminescent sensor Aqueous solution 21 – (Cen et al., 2020)
4 Cd-MOF-based luminescent sensor Aqueous solution 0.3 0–16 (Lv et al., 2018)
5 [ZnL]⋅2 H2O-based luminescent sensor Aqueous solution 0.92 1–1000 (Li et al., 2018)
[B] Bi3þ ions
6 CAU-1-(OH)2-based luminescent sensor Aqueous solution 2.16 0–100 (Gao et al., 2018)
7 N-hydroxy 1,8-naphthalimide luminescent sensor Water 2.78 – (Ramasamy and
Thambusamy, 2017)
8 Rhodamine 6 G-based luminescent sensor Drugs 2.69 10–35 (Zhang et al., 2018a)
9 Glutathione/copper NPs luminescent sensor – 10000 (0–100) × 103 (Wu et al., 2022)
[C] Pb2þ ions
10 MIL-53(Al)@CdTe-PEI electrochemiluminescent sensor Aqueous solution 2.4 × 10− 5 10− 4
- 10 (Feng et al., 2020)
11 Glutathione/thioglycolic acid/CdTe – 0.26 × 10− 3 – (Wang et al., 2012)
electrochemiluminescent sensor
12 CdS QDs electrochemiluminescent sensor Water 3.74 × 10− 5 (0.2–50) × 10− 3
(Xiao-Ping et al., 2014)
13 CdS QDs/graphene/gold electrochemiluminescent sensor Water 1 × 10− 4 (0.1–10) × 10− 3
(Lu et al., 2016)
14 CdSe/ZnS electrochemiluminescent sensor Aqueous solution 0.99 × 10¡3 (0–76) × 10¡3 (Wang et al., 2015b)
[D] Pd2þ ions
15 CAU-10-V-H-based luminescent sensor Water 0.11 – (Nandi et al., 2020)
16 {[Zn-(ATA)(L)]⋅H2O}n-based luminescent sensor Water 0.2 0–25 (Parmar et al., 2017)
17 3-(1-isoquinolinyl)imidazo[5, 1-a]isoquinoline luminescent Water 0.21 – (Mahata et al., 2020)
sensor
18 Quinoline-benzimidazole conjugate material luminescent Water 0.26 0.5–10 (Wang et al., 2021)
sensor
[E] Hg2þ ions
19 MIL-53(Al)@CdTe-PEI electrochemiluminescent sensor Aqueous solution 4.1 × 10− 6 10− 5 - 10− 1 (Feng et al., 2020)
20 Fe-single atom catalysts electrochemiluminescent sensor Aqueous solution 0.13 × 10− 3 (0.01 − 0.5) (Bushira et al., 2021)
× 10− 3
3
21 Polyluminol gold nanocomposite electrochemiluminescent Water 0.1 × 10− (0.3–200) (Raju and Kumar, 2021)
sensor × 10− 3
6
22 Gold NP-modified indium tin oxide electrochemiluminescent Water 5.1 × 10− (0.02–30) (Huang et al., 2015)
sensor × 10− 3
3
23 Micro fluidic paper-based electrochemiluminescent sensor Water 0.2 × 10− 0.0005–1 (Zhang et al., 2013)
2+
[F] Cu ions
− 3
24 NH2-MIL-53(Al)/Ppy-based electrochemical sensor Aqueous solution 3.84 × 10 0.015–6.294 (Wang et al., 2020b)
25 NH2-Ni-MOF-baed electrochemical sensor Water 6.3 × 10− 3 0.01–2 (Wan et al., 2021)
26 Graphene aerogel/UiO-66-NH2-based electrochemical sensor Water 8 × 10− 3 0.1–3.5 (Lu et al., 2018)
27 Cu3(HHTP)2-based electrochemical sensor Aqueous solution 0.331 100–10000 (Xu et al., 2022)
28 COFBTLP-based electrochemical sensor Water 18.6 × 10− 3 0.0564 –18 (Han et al., 2020)
29 Covalent-organic framework-based electrochemical sensor Water 1.14 × 10− 3 0.0034–4 (Pei et al., 2022)
[G] Fe2+ ions
30 MOF-253-based luminescent sensor Aqueous solution 0.5 5–100 (Lu et al., 2014)
31 [Eu2(IMBA)6(dmp)2]-based luminescent sensor Aqueous solution 1 20–80 (Zhao et al., 2018)
32 Ni-MOF based luminescent sensor Aqueous solution 25.03 – (Asiwal et al., 2022)
33 Triphenylamine/2,2′ -bipyridine-5,5′ -diformaldehyde Water 5.37 0–200 (Zhang et al., 2022b)
based luminescent sensor
34 Sicyanomethylene-4 H-pyran based luminescent sensor Living cells 4.5 0–100 (Yang et al., 2019)
[H] Agþ ions
35 Sm3þ@MIL-121-based luminescent sensor Aqueous solution 0.09 0–500 (Hao and Yan, 2015)
36 Carbon dot luminescent sensor Aqueous solution 0.386 (0–2.66)× 103 (Algarra et al., 2014)
37 [ZnIJBTA)2]n luminescent sensor Aqueous solution 9 – (Shao et al., 2022)
38 Bis-heteroleptic Ru(II) complex luminescent sensor Carbonate-bicarbonate buffer/ 0.46 0–250 (Sen et al., 2021)
CH3CN solution
39 {[Zn6Cl6(2,2′-dbpt)3]⋅4.5 H2O}n luminescent sensor Aqueous solution 6.4 102–182 (Cao et al., 2020)

20
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

complex formation between Pd2+ ions and vinyl functionality of sonochemical, microwave-assisted method, electrochemical, and
CAU-10-V-H materials (Nandi et al., 2020). More efforts are thus needed mechanochemical method) in relation to their merits and disadvantages.
to improve its practicality. The architectural versatility of Al-MOFs enables their facile functional­
ization with advanced sensing elements (e.g., fluorescent organic li­
4.2. Comparing the performance of Al-MOF-based sensors with other gands, lanthanide ions, and CdTe quantum dots). As such, the
common sensors functionalized Al-MOFs have become a favorable medium for the
detection of metal ions in the aqueous system using (i) fluorescence
The performance metrics (e.g., LOD and dynamic detection range) of intensity, (ii) quantum yield, (iii) conductivity, and (iv) electric catalytic
Al-MOF-based sensors were also compared with other common sensors activity. The analytical performance of various Al-MOF-based sensors
including MOF-based sensors and non-MOF-based sensors to learn more used for the detection of metal ions has been discussed to ensure their
about the uses of Al-MOF-based sensors in the detection of various metal superior sensing performance (in terms of detection limit) over non-Al-
ions (Table 4). For this purpose, the best Al-MOF-based sensing systems MOF-based sensors. However, some roadblocks still need to be
were compared with the best non-Al-MOF-based sensing systems for addressed in real-world applications as described below:
detecting various metal ions (e.g., Ag+, Pd2+, and Fe3+). As summarized
in Table 4, the aluminum CAU-1-(OH)2-based luminescent sensor (with (1) Extensive efforts should be put forward to guarantee the feasi­
LOD value of 2.16 µM), having a wide linear detection range bility of cyclic utilization of Al-MOF-based metal ion for their
(0–100 µM), outperformed the non-Al-MOF-based sensor (e.g., rhoda­ practical applications in the sensing fields. Moreover, the utili­
mine 6 G-based luminescent sensor, LOD: 2.69 µM, linear detection zation of some novel techniques (e.g., MOFs test paper) could be
range: 10–35 µM) for the detection of Bi3+ ions (Gao et al., 2018; Zhang considered as an effective tool to avoid possible lost issue of
et al., 2018a). In case of Fe3+ detection, an Al-MIL-53-N3-based lumi­ powder state MOFs during application. As such, more efficient
nescent sensor with an LOD value of 0.03 µM exhibited a higher sensing strategies should be proposed in the near future research.
ability than Cd-MOF-based luminescent sensor (LOD = 0.3 µM) (Das (2) More efforts should be made to explore continuous flow synthesis
et al., 2018; Lv et al., 2018). In addition, the Al-MOF-253-based lumi­ routes for maintaining the high quality (e.g., crystallinity and
nescent sensor exhibited significantly higher sensitivity for detecting narrow size distribution) of the fabricated Al-MOFs products. In
Fe2+ ions (0.5 µM) than a [Eu2(IMBA)6(dmp)2] MOF-based luminescent addition, continuous synthesis routes can be used to overcome
sensor (1 µM) (Lu et al., 2014; Zhao et al., 2018). For the detection of the shortcomings of the batch process (e.g., low efficiency and
Pd2+, the Al-CAU-10-V-H-based luminescent sensor offered a low LOD lack of flexibility). As such, a continuous flow synthesis system
value (0.11 µM) as compared to a non-Al-based zinc {[Zn-(ATA)(L)]⋅ can be employed to scale up their industrial applications.
H2O}n luminescent sensor (0.2 µM) (Nandi et al., 2020; Parmar et al., (3) The introduction of functional materials with a variety of prop­
2017). The detection limit of MIL-53(Al)@CdTe-PEI-based electro­ erties (e.g., excellent conductivity) into Al-MOFs can endow the
chemiluminescent sensor was noticeably lower for Pb2+ ions (2.4 ×10− 5 Al-MOF composites with upgraded properties for sensing appli­
µM) than that of a CdS QDs electrochemiluminescent sensor cations. As such, Al-MOF composites can make use of various
(3.74 ×10− 5 µM) (Feng et al., 2020; Xiao-Ping et al., 2014). Regarding sensitive signal transduction principles (e.g., electrochemical) for
the detection of monovalent ions (e.g., Ag+), a Sm3+@MIL-121-based highly sensitive and selective detection of analytes.
luminescent sensor showed a superior LOD value (0.09 µM) compared (4) Efforts should be put to upscale from the laboratory research into
with that of a carbon dot luminescent sensor (0.386 µM) (Hao and Yan, industrial applications. To date, the potential of Al-MOF-based
2015; Algarra et al., 2014). Nevertheless, in the Cu2+ sensing systems, sensors has mainly been explored at lab-scale. Thus, it is indis­
the covalent-organic framework (COF)-based electrochemical sensor (i. pensable to find a large-scale synthesis route to expand the real-
e., non-Al-MOF-based sensor) exhibited better sensitivity (LOD: world applications of Al-MOF-based sensors.
1.14 ×10− 3 µM) than those of NH2-MIL-53(Al)/Ppy-based electro­
chemical sensors (LOD: 3.84 ×10− 3 µM) (Wang et al., 2020b; Pei et al., Environmental implication
2022). The better performance of a COF-based electrochemical sensor
may be attributed to the abundant actives sites (e.g., N-S-N and –C–– N) Toxic heavy metals in the environment are responsible for the
for capturing Cu2+ ions in the highly ordered COF-based conjugated various diseases and disasters. They cause many environmental prob­
structure (Pei et al., 2022). lems. Therefore, their detection from environment is very important.
In light of these above findings, the sensing performance of Al-MOF- Chemical sensing detection of metal ions from water is regarded as high
based sensors showed a better performance in terms of lowest LOD value sensitivity and facile technology. Al-MOFs have received considerable
compared with other common MOF-based and non-MOF based sensors attention due to their merits of low cost, good water stability and
for various metal ions (e.g., Pb2+, Ag+, Bi3+, Pd2+, Fe3+, and Fe2+) as thermal stability. Al-MOFs sensing system can be formulated via direct
listed in Table 4. Moreover, the LOD of MIL-53(Al)@CdTe-PEI-based synthesis, post-synthetic modification, and composite material prepa­
electrochemiluminescent sensor was moderately superior in metal ions ration. Such formulation can efficiently enhance the interactions be­
detection (Pb2+: 2.4 ×10− 5 µM) to some reported classical techniques tween the Al-MOFs and target heavy metal ions to promote the selective
including ICP-OES (LOD: 4.83 ×10− 5 µM) and ICP-MS (LOD: detection of heavy metal ions. The development of Al-MOFs for highly
2.89 ×10− 5 µM) (Feng et al., 2020; Hosseinzadegan et al., 2019; Voica detection of heavy metal ions in liquid phase with respect to their per­
et al., 2009). As such, it is revealed that Al-MOF-based sensors can be formance and associated detection mechanisms.
regarded as a promising option for the detection of various metal ions in
aqueous system.
Declaration of Competing Interest
5. Conclusion and outlook
The authors declare that they have no known competing financial
Al-MOFs can overcome many disadvantages of common MOFs (e.g., interests or personal relationships that could have appeared to influence
poor thermal, chemical, and water stability) to allow the expansion of the work reported in this paper.
their applications toward various fields (e.g., detection of metals in
aqueous medium). In light of the beneficial properties of Al-MOFs, this Data Availability
comprehensive review is offered to describe the latest progress achieved
in their synthesis (e.g., solvo/hydrothermal, reflux method, Data will be made available on request.

21
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Acknowledgments framework for multi-purpose water-sorption-driven heat allocations. Nat. Commun.


11, 1–8.
Cohen, S.M., 2012. Postsynthetic methods for the functionalization of metal–organic
This work was supported partially by a grant from the National frameworks. Chem. Rev. 112, 970–1000.
Research Foundation of Korea (NRF) funded by the Ministry of Science Comotti, A., Bracco, S., Sozzani, P., Horike, S., Matsuda, R., Chen, J., Takata, M.,
and ITC (MSIT) of the Korean government (Grant No: Kubota, Y., Kitagawa, S., 2008. Nanochannels of two distinct cross-sections in a
porous Al-based coordination polymer. J. Am. Chem. Soc. 130, 13664–13672.
2021R1A3B1068304). Crawford, D., Casaban, J., Haydon, R., Giri, N., McNally, T., James, S.L., 2015. Synthesis
by extrusion: continuous, large-scale preparation of MOFs using little or no solvent.
References Chem. Sci. 6, 1645–1649.
Crawford, D.E., Casaban, J., 2016. Recent developments in mechanochemical materials
synthesis by extrusion. Adv. Mater. 28, 5747–5754.
Aguilera-Sigalat, J., Bradshaw, D., 2016. Synthesis and applications of metal-organic Cui, Y., Yue, Y., Qian, G., Chen, B., 2012. Luminescent functional metal–organic
framework–quantum dot (QD@ MOF) composites. Coord. Chem. Rev. 307, 267–291. frameworks. Chem. Rev. 112, 1126–1162.
Aguirre-Díaz, L.M., Gá ndara, F., Iglesias, M., Snejko, N., Gutiérrez-Puebla, E., Cui, Y., Chen, B., Qian, G., 2014. Lanthanide metal-organic frameworks for luminescent
Monge, M.Á., 2015. Tunable catalytic activity of solid solution metal–organic sensing and light-emitting applications. Coord. Chem. Rev. 273, 76–86.
frameworks in one-pot multicomponent reactions. J. Am. Chem. Soc. 137, Dao, X., Ni, Y., 2017. Al-Based coordination polymer nanotubes: simple preparation,
6132–6135. post-modification and application in Fe 3+ ions sensing. Dalton Trans. 46,
Aguirre-Díaz, L.M., Reinares-Fisac, D., Iglesias, M., Gutierrez-Puebla, E., Gandara, F., 5373–5383.
Snejko, N., Monge, M.Á., 2017. Group 13th metal-organic frameworks and their role Dao, X., Ni, Y., Pan, H., 2018. MIL-53 (Al)/Eu3+ luminescent nanocrystals: solvent-
in heterogeneous catalysis. Coord. Chem. Rev. 335, 1–27. adjusted shape-controllable synthesis and highly selective detections for Fe3+ ions,
Ahadi, N., Askari, S., Fouladitajar, A., Akbari, I., 2022. Facile synthesis of hierarchically Cr2O72− anions and acetone. Sens. Actuators B Chem. 271, 33–43.
structured MIL-53 (Al) with superior properties using an environmentally-friendly Das, A., Banesh, S., Trivedi, V., Biswas, S., 2018. Extraordinary sensitivity for H 2 S and
ultrasonic method for separating lead ions from aqueous solutions. Sci. Rep. 12, Fe (III) sensing in aqueous medium by Al-MIL-53-N 3 metal–organic framework: In
1–17. vitro and in vivo applications of H 2 S sensing. Dalton Trans. 47, 2690–2700.
Ahmed, S., Khan, F.S.A., Mubarak, N.M., Khalid, M., Tan, Y.H., Mazari, S.A., Karri, R.R., Devic, T., Serre, C., 2014. High valence 3p and transition metal based MOFs. Chem. Soc.
Abdullah, E.C., 2021. Emerging pollutants and their removal using visible-light Rev. 43, 6097–6115.
responsive photocatalysis–a comprehensive review. J. Environ. Chem. Eng. 9, Díaz-Ramírez, M.L., Vargas, B., Álvarez, J.R., Landeros-Rivera, B., Rivera-Almazo, M.,
106643. Ramos, C., Flores, J.G., Morales, E., Vargas, R., Garza, J., 2020. Fluorometric
Ahnfeldt, T., Stock, N., 2012. Synthesis of isoreticular CAU-1 compounds: effects of detection of iodine by MIL-53 (Al)-TDC. Dalton Trans. 49, 6572–6577.
linker and heating methods on the kinetics of the synthesis. CrystEngComm 14, Ding, M., Cai, X., Jiang, H.-L., 2019. Improving MOF stability: approaches and
505–511. applications. Chem. Sci. 10, 10209–10230.
Ahnfeldt, T., Guillou, N., Gunzelmann, D., Margiolaki, I., Loiseau, T., Férey, G., Duan, D., Si, X., Ding, Y., Li, L., Ma, G., Zhang, L., Jian, B., 2019. A novel molecularly
Senker, J., Stock, N., 2009. [Al4 (OH) 2 (OCH3) 4 (H2N-bdc) 3]⋅ x H2O: a imprinted electrochemical sensor based on double sensitization by MOF/CNTs and
12–connected porous metal–organic framework with an unprecedented aluminum- Prussian blue for detection of 17β-estradiol. Bioelectrochemistry 129, 211–217.
containing brick. Angew. Chem. Int. Ed. 48, 5163–5166. Fan, Y., Li, C., Zhang, X., Yang, X., Su, X., Ye, H., Li, N., 2019. Tröger’s base mixed matrix
Ahnfeldt, T., Moellmer, J., Guillerm, V., Staudt, R., Serre, C., Stock, N., 2011. High- membranes for gas separation incorporating NH2-MIL-53 (Al) nanocrystals.
throughput and time-resolved energy-dispersive X-ray diffraction (EDXRD) study of J. Membr. Sci. 573, 359–369.
the formation of CAU-1–(OH) 2: microwave and conventional heating. Chem. Eur. J. Fang, X., Zong, B., Mao, S., 2018. Metal–organic framework-based sensors for
17, 6462–6468. environmental contaminant sensing. Nano-Micro Lett. 10, 1–19.
Al-Attri, R., Halladj, R., Askari, S., 2022. Green route of flexible Al-MOF synthesis with Feng, D., Li, P., Tan, X., Wu, Y., Wei, F., Du, F., Ai, C., Luo, Y., Chen, Q., Han, H., 2020.
superior properties at low energy consumption assisted by ultrasound waves. Solid Electrochemiluminescence aptasensor for multiple determination of Hg2+ and Pb2+
State Sci. 123, 106782. ions by using the MIL-53 (Al)@ CdTe-PEI modified electrode. Anal. Chim. Acta 1100,
Algarra, M., Campos, B.B., Radotić, K., Mutavdžić, D., Bandosz, T., Jimenez-Jimenez, J., 232–239.
Rodriguez-Castellón, E., da Silva, J.C.E., 2014. Luminescent carbon nanoparticles: Feng, D., Wei, F., Wu, Y., Tan, X., Li, F., Lu, Y., Fan, G., Han, H., 2021. A novel signal
effects of chemical functionalization, and evaluation of Ag+ sensing properties. amplified electrochemiluminescence biosensor based on MIL-53 (Al)@ CdS QDs and
J. Mater. Chem. A 2, 8342–8351. SiO 2@ AuNPs for trichlorfon detection. Analyst 146, 1295–1302.
Asiwal, E.P., Shelar, D.S., Manjare, S.T., Pawar, S.D., 2022. Ni-MOF based luminescent Feng, J., Zhang, H., 2013. Hybrid materials based on lanthanide organic complexes: a
sensor for selective and rapid sensing of Fe (II) and Fe (III) ions. New J. Chem. review. Chem. Soc. Rev. 42, 387–410.
Banica, F.-G., 2012. Chemical Sensors and Biosensors: Fundamentals and Applications. Fröhlich, D., Pantatosaki, E., Kolokathis, P.D., Markey, K., Reinsch, H., Baumgartner, M.,
John Wiley & Sons. van der Veen, M.A., De Vos, D.E., Stock, N., Papadopoulos, G.K., 2016. Water
Barth, B., Mendt, M., Pöppl, A., Hartmann, M., 2015. Adsorption of nitric oxide in metal- adsorption behaviour of CAU-10-H: a thorough investigation of its
organic frameworks: Low temperature IR and EPR spectroscopic evaluation of the structure–property relationships. J. Mater. Chem. A 4, 11859–11869.
role of open metal sites. Microporous Mesoporous Mater. 216, 97–110. Gaab, M., Trukhan, N., Maurer, S., Gummaraju, R., Müller, U., 2012. The progression of
Bloch, E.D., Britt, D., Lee, C., Doonan, C.J., Uribe-Romo, F.J., Furukawa, H., Long, J.R., Al-based metal-organic frameworks–From academic research to industrial
Yaghi, O.M., 2010. Metal insertion in a microporous metal− organic framework production and applications. Microporous Mesoporous Mater. 157, 131–136.
lined with 2, 2′ -bipyridine. J. Am. Chem. Soc. 132, 14382–14384. Gándara, F., Furukawa, H., Lee, S., Yaghi, O.M., 2014. High methane storage capacity in
Bushira, F.A., Kitte, S.A., Xu, C., Li, H., Zheng, L., Wang, P., Jin, Y., 2021. Two- aluminum metal–organic frameworks. J. Am. Chem. Soc. 136, 5271–5274.
dimensional-plasmon-boosted iron single-atom electrochemiluminescence for the Gao, X., Zhao, H., Zhao, X., Li, Z., Gao, Z., Wang, Y., Huang, H., 2018. Aqueous phase
ultrasensitive detection of dopamine, hemin, and mercury. Anal. Chem. 93, sensing of bismuth ion using fluorescent metal-organic framework. Sens. Actuators B
9949–9957. Chem. 266, 323–328.
Cadiau, A., Lee, J.S., Damasceno Borges, D., Fabry, P., Devic, T., Wharmby, M.T., García Márquez, A., Demessence, A., Platero-Prats, A.E., Heurtaux, D., Horcajada, P.,
Martineau, C., Foucher, D., Taulelle, F., Jun, C.H., 2015. Design of hydrophilic metal Serre, C., Chang, J.S., Férey, G., de la Peña-O’Shea, V.A., Boissière, C., 2012. Green
organic framework water adsorbents for heat reallocation. Adv. Mater. 27, microwave synthesis of MIL-100 (Al, Cr, Fe) nanoparticles for thin-film elaboration.
4775–4780. Eur. J. Inorg. Chem. 2012, 5165–5174.
Camilli, L., Passacantando, M., 2018. Advances on sensors based on carbon nanotubes. Ghoorchian, A., Afkhami, A., Madrakian, T., Ahmadi, M., 2020. Chapter 9 -
Chemosensors 6, 62. electrochemical synthesis of MOFs. In: Mozafari, M. (Ed.), Metal-Organic
Cao, Y., Zhang, Y., Gu, L.-W., Qin, X.-M., Li, H.-Y., Bian, H.-D., Huang, F.-P., 2020. A zinc Frameworks for Biomedical Applications. Woodhead Publishing, pp. 177–195.
2+-dpbt framework: luminescence sensing of Cu 2+, Ag+, MnO 4− and Cr (VI)(Cr 2 Ghosh, S., Steinke, F., Rana, A., Alam, M., Biswas, S., Metal-Organic, A., 2021.
O 7 2− and CrO 4 2− ) ions. N. J. Chem. 44, 10681–10688. Framework with allyloxy functionalization for aqueous-phase fluorescence
Cen, P., Liang, C., Duan, L., Wang, M., Tian, D., Liu, X., 2020. A robust 3D In–MOF with recognition of Pd (II) Ion. Eur. J. Inorg. Chem. 2021, 3846–3851.
an imidazole acid ligand as a fluorescent sensor for sensitive and selective detection Guo, H., Wu, N., Xue, R., Liu, H., Li, L., Wang, M.-Y., Yao, W.-Q., Li, Q., Yang, W., 2020.
of Fe 3+ ions. N. J. Chem. 44, 16076–16081. Multifunctional Ln-MOF luminescent probe displaying superior capabilities for
Chen, D., Zhao, J., Zhang, P., Dai, S., 2019b. Mechanochemical synthesis of highly selective sensing of Fe3+ and Al3+ ions and nitrotoluene. Colloids Surf. A:
metal–organic frameworks. Polyhedron 162, 59–64. Physicochem. Eng. Asp. 585, 124094.
Chen, L., Liu, D., Peng, J., Du, Q., He, H., 2020. Ratiometric fluorescence sensing of Guo, Y., Zhang, J., Dong, L.Z., Xu, Y., Han, W., Fang, M., Liu, H.K., Wu, Y., Lan, Y.Q.,
metal-organic frameworks: tactics and perspectives. Coord. Chem. Rev. 404, 2017. Syntheses of exceptionally stable aluminum (III) metal–organic frameworks:
213113. how to grow high-quality, large, single crystals. Chem. Eur. J. 23, 15518–15528.
Chen, Z., Hanna, S.L., Redfern, L.R., Alezi, D., Islamoglu, T., Farha, O.K., 2019a. Halis, S., Reimer, N., Klinkebiel, A., Lüning, U., Stock, N., 2015. Four new Al-based
Reticular chemistry in the rational synthesis of functional zirconium cluster-based microporous metal-organic framework compounds with MIL-53-type structure
MOFs. Coord. Chem. Rev. 386, 32–49. containing functionalized extended linker molecules. Microporous Mesoporous
Cheng, X., Zhang, A., Hou, K., Liu, M., Wang, Y., Song, C., Zhang, G., Guo, X., 2013. Size- Mater. 216, 13–19.
and morphology-controlled NH 2-MIL-53 (Al) prepared in DMF–water mixed Han, J., Yu, J., Guo, Y., Wang, L., Song, Y., 2020. COFBTLP-1/three-dimensional
solvents. Dalton Trans. 42, 13698–13705. macroporous carbon electrode for simultaneous electrochemical detection of Cd2+,
Cho, K.H., Borges, D.D., Lee, U., Lee, J.S., Yoon, J.W., Cho, S.J., Park, J., Lombardo, W., Pb2+, Cu2+ and Hg2+. Sens. Actuators B: Chem. 321, 128498.
Moon, D., Sapienza, A., 2020. Rational design of a robust aluminum metal-organic

22
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Hao, J.-N., Yan, B., 2014. Highly sensitive and selective fluorescent probe for Ag+ based Li, C., Zhu, L., Yang, W., He, X., Zhao, S., Zhang, X., Tang, W., Wang, J., Yue, T., Li, Z.,
on a Eu 3+ post-functionalized metal–organic framework in aqueous media. 2019. Amino-functionalized Al–MOF for fluorescent detection of tetracyclines in
J. Mater. Chem. A 2, 18018–18025. milk. J. Agric. Food Chem. 67, 1277–1283.
Hao, J.-N., Yan, B., 2015. Ag+-sensitized lanthanide luminescence in Ln 3+ post- Li, H.-Y., Zhao, S.-N., Zang, S.-Q., Li, J., 2020a. Functional metal–organic frameworks as
functionalized metal–organic frameworks and Ag+ sensing. J. Mater. Chem. A 3, effective sensors of gases and volatile compounds. Chem. Soc. Rev. 49, 6364–6401.
4788–4792. Li, J., Hurlock, M.J., Goncharov, V.G., Li, X., Guo, X., Zhang, Q., 2021. Solvent-free and
Hao, Z., Song, X., Zhu, M., Meng, X., Zhao, S., Su, S., Yang, W., Song, S., Zhang, H., 2013. phase-selective synthesis of aluminum trimesate metal–organic frameworks. Inorg.
One-dimensional channel-structured Eu-MOF for sensing small organic molecules Chem. 60, 4623–4632.
and Cu 2+ ion. J. Mater. Chem. A 1, 11043–11050. Li, L., Chen, Y., Zhu, J.-J., 2017. Recent advances in electrochemiluminescence analysis.
Haque, E., Lo, V., Minett, A.I., Harris, A.T., Church, T.L., 2014. Dichotomous adsorption Anal. Chem. 89, 358–371.
behaviour of dyes on an amino-functionalised metal–organic framework, amino- Li, P., Zhou, L.-J., Yang, N.-N., Sui, Q., Gong, T., Gao, E.-Q., 2018. Metal–organic
MIL-101 (Al). J. Mater. Chem. A 2, 193–203. frameworks with extended viologen units: metal-dependent photochromism,
Henry, B., Samokhvalov, A., 2022. Hygroscopic metal-organic framework MIL-160 (Al): photomodulable fluorescence, and sensing properties. Cryst. Growth Des. 18,
in-situ time-dependent ATR-FTIR and gravimetric study of mechanism and kinetics 7191–7198.
of water vapor sorption. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 267, Li, S., Tan, L., Meng, X., 2020b. Nanoscale metal-organic frameworks: synthesis,
120550. biocompatibility, imaging applications, and thermal and dynamic therapy of tumors.
Hosseinzadegan, S., Nischkauer, W., Bica, K., Limbeck, A., 2019. FI-ICP-OES Adv. Funct. Mater. 30, 1908924.
determination of Pb in drinking water after pre-concentration using magnetic Li, Z., Wu, Yn, Li, J., Zhang, Y., Zou, X., Li, F., 2015. The metal–organic framework MIL-
nanoparticles coated with ionic liquid. Microchem. J. 146, 339–344. 53 (Al) constructed from multiple metal sources: alumina, aluminum hydroxide, and
Hu, Z., Deibert, B.J., Li, J., 2014. Luminescent metal–organic frameworks for chemical boehmite. Chem. Eur. J. 21, 6913–6920.
sensing and explosive detection. Chem. Soc. Rev. 43, 5815–5840. Lin, C., Zou, Z., Lei, Z., Wang, L., Song, Y., 2020. Fluorescent metal-organic frameworks
Hua, Y., Kukkar, D., Brown, R.J., Kim, K.-H., 2022. Recent advances in the synthesis of MIL-101 (Al)-NH2 for rapid and sensitive detection of ellagic acid. Spectrochim. Acta
and sensing applications for metal-organic framework-molecularly imprinted Part A: Mol. Biomol. Spectrosc. 242, 118739.
polymer (MOF-MIP) composites. Crit. Rev. Environ. Sci. Technol. 1–32. Liu, C., Yan, B., 2016. A novel photofunctional hybrid material of pyrene functionalized
Huang, R.-F., Liu, H.-X., Gai, Q.-Q., Liu, G.-J., Wei, Z., 2015. A facile and sensitive metal-organic framework with conformation change for fluorescence sensing of Cu2
electrochemiluminescence biosensor for Hg2+ analysis based on a dual-function +. Sens. Actuators B: Chem. 235, 541–546.
oligonucleotide probe. Biosens. Bioelectron. 71, 194–199. Liu, D., Lu, K., Poon, C., Lin, W., 2014. Metal–organic frameworks as sensory materials
Huangfu, M., Wang, M., Lin, C., Wang, J., Wu, P., 2021. Luminescent metal–organic and imaging agents. Inorg. Chem. 53, 1916–1924.
frameworks as chemical sensors based on “mechanism–response”: a review. Dalton Liu, Y., Pan, M., Yang, Q.-Y., Fu, L., Li, K., Wei, S.-C., Su, C.-Y., 2012. Dual-emission from
Trans. 50, 3429–3449. a single-phase Eu–Ag metal–organic framework: an alternative way to get white-
Isaeva, V., Tarasov, A., Starannikova, L., Yampol’skii, Y.P., Alent’ev, A.Y., Kustov, L., light phosphor. Chem. Mater. 24, 1954–1960.
2015. Microwave-assisted synthesis of mesoporous metal-organic framework Liu, Y., Xie, X.-Y., Cheng, C., Shao, Z.-S., Wang, H.-S., 2019. Strategies to fabricate
NH2—MIL-101 (Al). Russ. Chem. Bull. 64, 2791–2795. metal–organic framework (MOF)-based luminescent sensing platforms. J. Mater.
Jayaramulu, K., Kanoo, P., George, S.J., Maji, T.K., 2010. Tunable emission from a Chem. C 7, 10743–10763.
porous metal–organic framework by employing an excited-state intramolecular Lo, S.-H., Chien, C.-H., Lai, Y.-L., Yang, C.-C., Lee, J.J., Raja, D.S., Lin, C.-H., 2013.
proton transfer responsive ligand. Chem. Commun. 46, 7906–7908. A mesoporous aluminium metal–organic framework with 3 nm open pores. J. Mater.
Jeremias, F., Fröhlich, D., Janiak, C., Henninger, S.K., 2014. Advancement of sorption- Chem. A 1, 324–329.
based heat transformation by a metal coating of highly-stable, hydrophilic Lodeiro, C., Capelo, J.L., Oliveira, E., Lodeiro, J.F., 2019. New Toxic Emerging
aluminium fumarate MOF. RSC Adv. 4, 24073–24082. Contaminants: Beyond the Toxicological Effects. Springer, pp. 1–4.
Johnson, M.L., Ackers, G.K., Holt, J.M., 1998. Energetics of Biological Macromolecules. Loiseau, T., Serre, C., Huguenard, C., Fink, G., Taulelle, F., Henry, M., Bataille, T.,
Elsevier. Férey, G., 2004. A rationale for the large breathing of the porous aluminum
Kalaj, M., Cohen, S.M., 2020. Postsynthetic modification: an enabling technology for the terephthalate (MIL-53) upon hydration. Chem. Eur. J. 10, 1373–1382.
advancement of metal–organic frameworks. ACS Cent. Sci. 6, 1046–1057. Loiseau, T., Mellot-Draznieks, C., Muguerra, H., Férey, G., Haouas, M., Taulelle, F., 2005.
Kalhor, S., Zarei, M., Zolfigol, M.A., Sepehrmansourie, H., Nematollahi, D., Alizadeh, S., Hydrothermal synthesis and crystal structure of a new three-dimensional aluminum-
Shi, H., Arjomandi, J., 2021. Anodic electrosynthesis of MIL-53 (Al)-N (CH2PO3H2) organic framework MIL-69 with 2, 6-naphthalenedicarboxylate (ndc), Al (OH)(ndc)⋅
2 as a mesoporous catalyst for synthesis of novel (N-methyl-pyrrol)-pyrazolo [3, 4-b] H2O. Comptes Rendus Chim. 8, 765–772.
pyridines via a cooperative vinylogous anomeric based oxidation. Sci. Rep. 11, 1–20. Loiseau, T., Lecroq, L., Volkringer, C., Marrot, J., Férey, G., Haouas, M., Taulelle, F.,
Kamel, R.M., Shahat, A., Anwar, Z.M., El-Kady, H.A., Kilany, E.M., 2021. A novel Bourrelly, S., Llewellyn, P.L., Latroche, M., 2006. MIL-96, a porous aluminum
sensitive and selective chemosensor for fluorescent detection of Zn 2+ in cosmetics trimesate 3D structure constructed from a hexagonal network of 18-membered rings
creams based on a covalent post functionalized Al-MOF. N. J. Chem. 45, 8054–8063. and μ 3-oxo-centered trinuclear units. J. Am. Chem. Soc. 128, 10223–10230.
Khan, N.A., Lee, J.S., Jeon, J., Jun, C.-H., Jhung, S.H., 2012. Phase-selective synthesis Loiseau, T., Volkringer, C., Haouas, M., Taulelle, F., Férey, G., 2015. Crystal chemistry of
and phase-conversion of porous aluminum-benzenetricarboxylates with microwave aluminium carboxylates: From molecular species towards porous infinite three-
irradiation. Microporous Mesoporous Mater. 152, 235–239. dimensional networks. Comptes Rendus Chim. 18, 1350–1369.
Khan, S., Guan, Q., Liu, Q., Qin, Z., Rasheed, B., Liang, X., Yang, X., 2021. Synthesis, Lu, L., Guo, L., Li, J., Kang, T., Cheng, S., 2016. Electrochemiluminescent detection of
modifications and applications of MILs metal-organic frameworks for environmental Pb2+ by graphene/gold nanoparticles and CdSe quantum dots. Appl. Surf. Sci. 388,
remediation: the cutting-edge review. Sci. Total Environ., 152279 431–436.
Kim, H.S., Yu, K., Puthiaraj, P., Ahn, W.-S., 2020. CO2 cycloaddition to epichlorohydrin Lu, M., Deng, Y., Luo, Y., Lv, J., Li, T., Xu, J., Chen, S.-W., Wang, J., 2018. Graphene
over an aluminum fumarate metal-organic framework synthesized by a aerogel–metal–organic framework-based electrochemical method for simultaneous
sonochemical route. Microporous Mesoporous Mater. 306, 110432. detection of multiple heavy-metal ions. Anal. Chem. 91, 888–895.
Kim, J., Kim, W.Y., Ahn, W.-S., 2012. Amine-functionalized MIL-53 (Al) for CO2/N2 Lu, Y., Yan, B., Liu, J.-L., 2014. Nanoscale metal–organic frameworks as highly sensitive
separation: effect of textural properties. Fuel 102, 574–579. luminescent sensors for Fe 2+ in aqueous solution and living cells. Chem. Commun.
Kim, S.-Y., Kang, J.H., Kim, S.-I., Bae, Y.-S., 2019. Extraordinarily large and stable 50, 9969–9972.
methane delivery of MIL-53 (Al) under LNG-ANG conditions. Chem. Eng. J. 365, Lu, Y., Liu, C., Mei, C., Sun, J., Lee, J., Wu, Q., Hubbe, M.A., Li, M.-C., 2022. Recent
242–248. advances in metal organic framework and cellulose nanomaterial composites. Coord.
Kreno, L.E., Leong, K., Farha, O.K., Allendorf, M., Van Duyne, R.P., Hupp, J.T., 2012. Chem. Rev. 461, 214496.
Metal–organic framework materials as chemical sensors. Chem. Rev. 112, Luo, J., Liu, B.-S., Zhang, X.-R., Liu, R.-T., 2019. A Eu3+ post-functionalized metal-
1105–1125. organic framework as fluorescent probe for highly selective sensing of Cu2+ in
Krüger, M., Siegel, R., Dreischarf, A., Reinsch, H., Senker, J., Stock, N., 2015. [Al2 (OH) 2 aqueous media. J. Mol. Struct. 1177, 444–448.
(TCPB)]–an Al-MOF based on a tetratopic linker molecule. Microporous Mesoporous Lustig, W.P., Mukherjee, S., Rudd, N.D., Desai, A.V., Li, J., Ghosh, S.K., 2017.
Mater. 216, 27–35. Metal–organic frameworks: functional luminescent and photonic materials for
Lee, S., Kapustin, E.A., Yaghi, O.M., 2016. Coordinative alignment of molecules in chiral sensing applications. Chem. Soc. Rev. 46, 3242–3285.
metal-organic frameworks. Science 353, 808–811. Lv, R., Chen, Z., Fu, X., Yang, B., Li, H., Su, J., Gu, W., Liu, X., 2018. A highly selective
Lee, Y.-R., Kim, J., Ahn, W.-S., 2013. Synthesis of metal-organic frameworks: a mini and fast-response fluorescent probe based on Cd-MOF for the visual detection of Al3
review. Korean J. Chem. Eng. 30, 1667–1680. + ion and quantitative detection of Fe3+ ion. J. Solid State Chem. 259, 67–72.
Lenzen, D., Bendix, P., Reinsch, H., Fröhlich, D., Kummer, H., Möllers, M., Hügenell, P.P., Ma, X., Zhang, X., Han, L., Hao, Z., Yong, S., Multi-response, A., 2021. Aluminum metal-
Gläser, R., Henninger, S., Stock, N., 2018. Scalable green synthesis and full-scale test organic frameworks for fluorescence sensing of Fe3+, Sr2+, SiO3 2− and toluene.
of the metal–organic framework CAU-10–H for use in adsorption-driven chillers. Methods Appl. Fluoresc. 9, 025007.
Adv. Mater. 30, 1705869. Mahata, S., Bhattacharya, A., Kumar, J.P., Mandal, B.B., Manivannan, V., 2020. Naked-
Lenzen, D., Zhao, J., Ernst, S.-J., Wahiduzzaman, M., Ken Inge, A., Fröhlich, D., Xu, H., eye detection of Pd2+ ion using a highly selective fluorescent heterocyclic probe by
Bart, H.-J., Janiak, C., Henninger, S., 2019. A metal–organic framework for efficient “turn-off” response and in-vitro live cell imaging. J. Photochem. Photobiol. A: Chem.
water-based ultra-low-temperature-driven cooling. Nat. Commun. 10, 1–9. 394, 112441.
Leubner, S., Stäglich, R., Franke, J., Jacobsen, J., Gosch, J., Siegel, R., Reinsch, H., Majhi, S.M., Ali, A., Rai, P., Greish, Y., Alzamly, A., Surya, S.G., Qamhieh, N.,
Maurin, G., Senker, J., Yot, P.G., 2020. Solvent impact on the properties of Mahmoud, S.T., 2022. Metal-organic frameworks for advanced transducers based
benchmark metal–organic frameworks: acetonitrile-based synthesis of CAU-10, Ce- gas sensors: review and perspectives. Nanoscale Adv.
UiO-66, and Al-MIL-53. Chem. Eur. J. 26, 3877–3883. Mandal, S., Natarajan, S., Mani, P., Pankajakshan, A., 2021. Post-synthetic modification
of metal–organic frameworks toward applications. Adv. Funct. Mater. 31, 2006291.

23
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Martinez Joaristi, A., Juan-Alcañiz, J., Serra-Crespo, P., Kapteijn, F., Gascon, J., 2012. Safaei, M., Foroughi, M.M., Ebrahimpoor, N., Jahani, S., Omidi, A., Khatami, M., 2019.
Electrochemical synthesis of some archetypical Zn2+, Cu2+, and Al3+ metal A review on metal-organic frameworks: synthesis and applications. TrAC Trends
organic frameworks. Cryst. Growth Des. 12, 3489–3498. Anal. Chem. 118, 401–425.
McLeod, C.L., Shaulis, B.J., 2018. Rare earth elements in planetary crusts: insights from Samokhvalov, A., 2018. Aluminum metal–organic frameworks for sorption in solution: a
chemically evolved igneous suites on Earth and the Moon. Minerals 8, 455. review. Coord. Chem. Rev. 374, 236–253.
Moreau, F., Marrot, J., Banse, F., Serre, C., Tissot, A., 2020. Sequential installation of Fe Schilling, L.H., Reinsch, H., Stock, N., 2016. Synthesis, structure, and selected properties
(ii) complexes in MOFs: towards the design of solvatochromic porous solids. of aluminum-, gallium-, and indium-based metal–organic frameworks, the chemistry
J. Mater. Chem. C. 8, 16826–16833. of metal–organic frameworks: synthesis. Charact., Appl. 1, 105–135.
Nandi, S., Mondal, A., Reinsch, H., Biswas, S., 2019. An ultra-robust luminescent CAU-10 Schlu¨sener, C., Xhinovci, M., Ernst, S.-J., Schmitz, A., Tannert, N., Janiak, C., 2019.
MOF acting as a fluorescent “turn-off” sensor for Cr2O72− in aqueous medium. Solid-solution mixed-linker synthesis of isoreticular Al-based MOFs for an easy
Inorg. Chim. Acta 497, 119078. hydrophilicity tuning in water-sorption heat transformations. Chem. Mater. 31,
Nandi, S., Reinsch, H., Biswas, S., 2020. A vinyl functionalized mixed linker CAU-10 4051–4062.
metal-organic framework acting as a fluorescent sensor for the selective detection of Schlüsener, C., Jordan, D.N., Xhinovci, M., Ntep, T.J.M.M., Schmitz, A., Giesen, B.,
H2S and palladium (II). Microporous Mesoporous Mater. 293, 109790. Janiak, C., 2020. Probing the limits of linker substitution in aluminum MOFs
Niekiel, F., Ackermann, M., Guerrier, P., Rothkirch, A., Stock, N., 2013. Aluminum-1, 4- through water vapor sorption studies: mixed-MOFs instead of mixed-linker CAU-23
cyclohexanedicarboxylates: High-throughput and temperature-dependent in situ and MIL-160 materials. Dalton Trans. 49, 7373–7383.
EDXRD studies. Inorg. Chem. 52, 8699–8705. Schneemann, A., Bon, V., Schwedler, I., Senkovska, I., Kaskel, S., Fischer, R.A., 2014.
Niekiel, F., Lannoeye, J., Reinsch, H., Munn, A.S., Heerwig, A., Zizak, I., Kaskel, S., Flexible metal–organic frameworks. Chem. Soc. Rev. 43, 6062–6096.
Walton, R.I., De Vos, D., Llewellyn, P., 2014. Conformation-controlled sorption Schroeder, V., Savagatrup, S., He, M., Lin, S., Swager, T.M., 2018. Carbon nanotube
properties and breathing of the aliphatic Al-MOF [Al (OH)(CDC)]. Inorg. Chem. 53, chemical sensors. Chem. Rev. 119, 599–663.
4610–4620. Sen, B., Kumar Patra, S., Rabha, M., Kumar Sheet, S., Aguan, K., Samanta, D., Khatua, S.,
Paixão, T.R.L.C., Reddy, S.M., 2017. Materials for Chemical Sensing. Springer. 2021. Luminescence detection of Ag+ and phosphate ions by a ruthenium (II)
Paolesse, R., Nardis, S., Monti, D., Stefanelli, M., Di Natale, C., 2017. Porphyrinoids for complex-based multianalyte probe: a combined spectroscopic, crystallographic, and
chemical sensor applications. Chem. Rev. 117, 2517–2583. theoretical approach. Eur. J. Inorg. Chem. 2021, 3549–3560.
Parmar, B., Rachuri, Y., Bisht, K.K., Suresh, E., 2017. Mixed-ligand LMOF fluorosensors Senkovska, I., Hoffmann, F., Fröba, M., Getzschmann, J., Böhlmann, W., Kaskel, S., 2009.
for detection of Cr (VI) oxyanions and Fe3+/Pd2+ cations in aqueous media. Inorg. New highly porous aluminium based metal-organic frameworks: Al (OH)(ndc)(ndc=
Chem. 56, 10939–10949. 2, 6-naphthalene dicarboxylate) and Al (OH)(bpdc)(bpdc= 4, 4′ -biphenyl
Pei, L., Su, J., Yang, H., Wu, Y., Du, Y., Zhu, Y., 2022. A novel covalent-organic dicarboxylate). Microporous Mesoporous Mater. 122, 93–98.
framework for highly sensitive detection of Cd2+, Pb2+, Cu2+ and Hg2. Seoane, B., Dikhtiarenko, A., Mayoral, A., Tellez, C., Coronas, J., Kapteijn, F., Gascon, J.,
Microporous Mesoporous Mater. 333, 111742. 2015. Metal organic framework synthesis in the presence of surfactants: towards
Pellow, R., Vala, M., 1989. The external heavy atom effEct: Theory of spin–orbit coupling hierarchical MOFs? CrystEngComm 17, 1693–1700.
of alkali and noble metals in rare gas matrices. J. Chem. Phys. 90, 5612–5621. Serra-Crespo, P., Ramos-Fernandez, E.V., Gascon, J., Kapteijn, F., 2011. Synthesis and
Pereira, L.C., de Souza, A.O., Bernardes, M.F.F., Pazin, M., Tasso, M.J., Pereira, P.H., characterization of an amino functionalized MIL-101 (Al): separation and catalytic
Dorta, D.J., 2015. A perspective on the potential risks of emerging contaminants to properties. Chem. Mater. 23, 2565–2572.
human and environmental health. Environ. Sci. Pollut. Res. 22, 13800–13823. Shao, J.-J., Ni, J.-L., Liang, Y., Li, G.-J., Chen, L.-Z., Wang, F.-M., 2022. Luminescent
Permyakova, A., Skrylnyk, O., Courbon, E., Affram, M., Wang, S., Lee, U.H., Valekar, A. MOFs for selective sensing of Ag+ and other ions (Fe (iii) and Cr (vi)) in aqueous
H., Nouar, F., Mouchaham, G., Devic, T., 2017. Synthesis optimization, shaping, and solution. CrystEngComm 24, 2479–2484.
heat reallocation evaluation of the hydrophilic metal–organic framework MIL-160 Shi, J., Han, R., Lu, S., Liu, Q., metal-OH, A., 2021a. group modification strategy to
(Al). ChemSusChem 10, 1419–1426. prepare highly-hydrophobic MIL-53-Al for efficient acetone capture under humid
Qin, Y.-y, Wang, Q.-y, Ge, J.-L., Wu, F., 2021. Microwave ultrasound-assisted synthesis of conditions. J. Environ. Sci. 107, 111–123.
NH2-MIL-53 (Al) for fluorescence detection of organosulfur compounds in model Shi, L., Li, N., Wang, D., Fan, M., Zhang, S., Gong, Z., 2021b. Environmental pollution
fuel. Inorg. Chem. Commun. 132, 108828. analysis based on the luminescent metal organic frameworks: a review. TrAC Trends
A. Rahmani, A. Shabanloo, S. Zabihollahi, M. Salari, M. Leili, M. Khazaei, S. Alizadeh, D. Anal. Chem. 134, 116131.
Nematollahi, Facile Fabrication of Amino-Functionalized MIL-68 (Al) Metal-organic Sikka, R., Kumar, P., Lee, J., Sonne, C., 2022. Aqueous-phase biofunctionalized NH2-
Framework for Effective Adsorption of Arsenate (As (V)), (2022). MIL-53 (Al) MOF for biosensing applications. J. Porous Mater. 1–8.
Raja, D.S., Chang, I.-H., Jiang, Y.-C., Chen, H.-T., Lin, C.-H., 2015. Enhanced gas sorption Stavitski, E., Goesten, M., Juan-Alcañiz, J., Martinez-Joaristi, A., Serra-Crespo, P.,
properties of a new sulfone functionalized aluminum metal-organic framework: Petukhov, A.V., Gascon, J., Kapteijn, F., 2011. Kinetic control of metal–organic
Synthesis, characterization, and DFT studies. Microporous Mesoporous Mater. 216, framework crystallization investigated by time-resolved in situ X-Ray scattering.
20–26. Angew. Chem. 123, 9798–9802.
Raju, C.V., Kumar, S.S., 2021. Co-reactant-free self-enhanced solid-state Steenhaut, T., Filinchuk, Y., Hermans, S., 2021. Aluminium-based MIL-100 (Al) and MIL-
electrochemiluminescence platform based on polyluminol-gold nanocomposite for 101 (Al) metal–organic frameworks, derivative materials and composites: synthesis,
signal-on detection of mercury ion. Sci. Rep. 11, 1–11. structure, properties and applications. J. Mater. Chem. A 9, 21483–21509.
Ramasamy, K., Thambusamy, S., 2017. Dual emission and pH based naphthalimide Stock, N., 2014. Synthesis and structures of aluminum-based metal-organic frameworks.
derivative fluorescent sensor for the detection of Bi3+. Sens. Actuators B: Chem. Metal. Org. Framew. Mater. 1–16.
247, 632–640. Stock, N., Biswas, S., 2012. Synthesis of metal-organic frameworks (MOFs): routes to
Ramírez-Malule, H., Quinones-Murillo, D.H., Manotas-Duque, D., 2020. Emerging various MOF topologies, morphologies, and composites. Chem. Rev. 112, 933–969.
contaminants as global environmental hazards. A bibliometric analysis. Emerg. Sud, D., Kaur, G., 2021. A comprehensive review on synthetic approaches for metal-
Contam. 6, 179–193. organic frameworks: From traditional solvothermal to greener protocols. Polyhedron
Raptopoulou, C.P., 2021. Metal-organic frameworks: Synthetic methods and potential 193, 114897.
applications. Materials 14, 310. Sun, L., Yin, M., Li, Z., Tang, S., 2022. Facile microwave-assisted solvothermal synthesis
Reimer, N., Reinsch, H., Inge, A.K., Stock, N., 2015. New Al-MOFs based on of rod-like aluminum terephthalate [MIL-53 (Al)] for CO2 adsorption. J. Ind. Eng.
sulfonyldibenzoate ions: a rare example of intralayer porosity. Inorg. Chem. 54, Chem.
492–501. Sun, Y., Zhou, H.-C., 2015. Recent progress in the synthesis of metal–organic
Reinsch, H., Feyand, M., Ahnfeldt, T., Stock, N., 2012a. CAU-3: A new family of porous frameworks. Sci. Technol. Adv. Mater.
MOFs with a novel Al-based brick:[Al 2 (OCH 3) 4 (O 2 CX-CO 2)](X= aryl). Dalton Taheri, A., Babakhani, E.G., Towfighi, J., 2018. Study of synthesis parameters of MIL-53
Trans. 41, 4164–4171. (Al) using experimental design methodology for CO2/CH4 separation. Adsorpt. Sci.
Reinsch, H., Marszałek, B., Wack, J., Senker, J., Gil, B., Stock, N., 2012b. A new Al-MOF Technol. 36, 247–269.
based on a unique column-shaped inorganic building unit exhibiting strongly Tannert, N., Jansen, C., Nießing, S., Janiak, C., 2019. Robust synthesis routes and
hydrophilic sorption behaviour. Chem. Commun. 48, 9486–9488. porosity of the Al-based metal–organic frameworks Al-fumarate, CAU-10-H and MIL-
Reinsch, H., Kru¨ger, M., Marrot, J., Stock, N., 2013a. First keto-functionalized 160. Dalton Trans. 48, 2967–2976.
microporous al-based metal–organic framework:[Al (OH)(O2C-C6H4-CO-C6H4- Tehrani, M.S., Zare-Dorabei, R., 2016a. Competitive removal of hazardous dyes from
CO2)]. Inorg. Chem. 52, 1854–1859. aqueous solution by MIL-68 (Al): derivative spectrophotometric method and
Reinsch, H., van der Veen, M.A., Gil, B., Marszalek, B., Verbiest, T., De Vos, D., Stock, N., response surface methodology approach. Spectrochim. Acta Part A Mol. Biomol.
2013b. Structures, sorption characteristics, and nonlinear optical properties of a new Spectrosc. 160, 8–18.
series of highly stable aluminum MOFs. Chem. Mater. 25, 17–26. Tehrani, M.S., Zare-Dorabei, R., 2016b. Highly efficient simultaneous ultrasonic-assisted
Reinsch, H., De Vos, D., Stock, N., 2013c. Structure and Properties of [Al4 (OH) 8 (o- adsorption of methylene blue and rhodamine B onto metal organic framework MIL-
C6H4 (CO2) 2) 2]⋅H2O, a layered aluminum phthalate. Z. Anorg. Allg. Chem. 639, 68 (Al): central composite design optimization. RSC Adv. 6, 27416–27425.
2785–2789. Thomas-Hillman, I., Stevens, L.A., Lange, M., Möllmer, J., Lewis, W., Dodds, C.,
Ren, M., Wang, H., Liu, Y., Ma, Q., Jia, W., Liu, M., Wang, H., Lu, Y., 2020. Fluorescent Kingman, S.W., Laybourn, A., 2019. Developing a sustainable route to
determination of mercury (II) and glutathione using amino-MIL-53 (Al) nanosheets. environmentally relevant metal–organic frameworks: ultra-rapid synthesis of MFM-
Anal. Lett. 53, 2700–2714. 300 (Al) using microwave heating. Green. Chem. 21, 5039–5045.
Rubio-Martinez, M., Avci-Camur, C., Thornton, A.W., Imaz, I., Maspoch, D., Hill, M.R., Tricoli, A., Righettoni, M., Teleki, A., 2010. Semiconductor gas sensors: dry synthesis and
2017. New synthetic routes towards MOF production at scale. Chem. Soc. Rev. 46, application. Angew. Chem. Int. Ed. 49, 7632–7659.
3453–3480. Vinu, M., Lin, W.-C., Senthil Raja, D., Han, J.-L., Lin, C.-H., 2017. Microwave-assisted
synthesis of nanoporous aluminum-based coordination polymers as catalysts for
selective sulfoxidation reaction. Polymers 9, 498.

24
Y. hua et al. Journal of Hazardous Materials 444 (2023) 130422

Voica, C., Dehelean, A., Pamula, A., 2009. Method validation for determination of heavy Xu, X.-Y., Yan, B., 2016. Fabrication and application of a ratiometric and colorimetric
metals in wine and slightly alcoholic beverages by ICP-MS. In: Journal of Physics: fluorescent probe for Hg 2+ based on dual-emissive metal–organic framework
Conference Series. IOP Publishing, 012036. hybrids with carbon dots and Eu 3. J. Mater. Chem. C 4, 1543–1549.
Volkringer, C., Popov, D., Loiseau, T., Guillou, N., Ferey, G., Haouas, M., Taulelle, F., Yan, B., 2017. Lanthanide-functionalized metal–organic framework hybrid systems to
Mellot-Draznieks, C., Burghammer, M., Riekel, C., 2007. A microdiffraction set-up create multiple luminescent centers for chemical sensing. Acc. Chem. Res. 50,
for nanoporous metal–organic-framework-type solids. Nat. Mater. 6, 760–764. 2789–2798.
Volkringer, C., Loiseau, T., Guillou, N., Férey, G., Haouas, M., Taulelle, F., Yang, C.-X., Ren, H.-B., Yan, X.-P., 2013. Fluorescent metal–organic framework MIL-53
Audebrand, N., Margiolaki, I., Popov, D., Burghammer, M., 2009a. Structural (Al) for highly selective and sensitive detection of Fe3+ in aqueous solution. Anal.
transitions and flexibility during dehydration− rehydration process in the MOF-type Chem. 85, 7441–7446.
aluminum pyromellitate Al2 (OH) 2 [C10O8H2](MIL-118). Cryst. Growth Des. 9, Yang, G.L., Jiang, X.L., Xu, H., Zhao, B., 2021. Applications of MOFs as luminescent
2927–2936. sensors for environmental pollutants. Small 17, 2005327.
Volkringer, C., Loiseau, T., Haouas, M., Taulelle, F., Popov, D., Burghammer, M., Yang, J., Wang, J., Deng, S., Li, J., 2016. Improved synthesis of trigone trimer cluster
Riekel, C., Zlotea, C., Cuevas, F., Latroche, M., 2009a. Occurrence of uncommon metal organic framework MIL-100Al by a later entry of methyl groups. Chem.
infinite chains consisting of edge-sharing octahedra in a porous metal organic Commun. 52, 725–728.
framework-type aluminum pyromellitate Al4 (OH) 8 [C10O8H2](MIL-120): Yang, Q., Vaesen, S., Vishnuvarthan, M., Ragon, F., Serre, C., Vimont, A., Daturi, M., De
synthesis, structure, and gas sorption properties. Chem. Mater. 21, 5783–5791. Weireld, G., Maurin, G., 2012b. Probing the adsorption performance of the hybrid
Volkringer, C., Loiseau, T., Guillou, N., Férey, G., Elkaïm, E., 2009b. Syntheses and porous MIL-68 (Al): a synergic combination of experimental and modelling tools.
structures of the MOF-type series of metal 1, 4, 5, 8,-naphthalenetetracarboxylates J. Mater. Chem. 22, 10210–10220.
M2 (OH) 2 [C14O8H4](Al, Ga, In) with infinite trans-connected M–OH–M chains Yang, S., Sun, J., Ramirez-Cuesta, A.J., Callear, S.K., David, W.I., Anderson, D.P.,
(MIL-122). Solid State Sci. 11, 1507–1512. Newby, R., Blake, A.J., Parker, J.E., Tang, C.C., 2012a. Selectivity and direct
Volkringer, C., Loiseau, T., Guillou, N., Ferey, G., Haouas, M., Taulelle, F., Elkaim, E., visualization of carbon dioxide and sulfur dioxide in a decorated porous host. Nat.
Stock, N., 2010a. High-throughput aided synthesis of the porous metal− organic Chem. 4, 887–894.
framework-type aluminum pyromellitate, MIL-121, with extra carboxylic acid Yang, X., Wang, Y., Liu, R., Zhang, Y., Tang, J., Yang, E.-B., Zhang, D., Zhao, Y., Ye, Y.,
functionalization. Inorg. Chem. 49, 9852–9862. 2019. A novel ICT-based two photon and NIR fluorescent probe for labile Fe2+
Volkringer, C., Loiseau, T., Devic, T., Férey, G., Popov, D., Burghammer, M., Riekel, C., detection and cell imaging in living cells. Sens. Actuators B Chem. 288, 217–224.
2010b. A layered coordination polymer based on an azodibenzoate linker connected Yin, H.-Q., Yang, J.-C., Yin, X.-B., 2017. Ratiometric fluorescence sensing and real-time
to aluminium (MIL-129). CrystEngComm 12, 3225–3228. detection of water in organic solvents with one-pot synthesis of Ru@ MIL-101 (Al)–
Volkringer, C., Loiseau, T., Guillou, N., Férey, G., Popov, D., Burghammer, M., Riekel, C., NH2. Anal. Chem. 89, 13434–13440.
2013. Synthesis and structural characterization of metal–organic frameworks with Yin, Z., Wan, S., Yang, J., Kurmoo, M., Zeng, M.-H., 2019. Recent advances in post-
the mellitate linker M2 (OH) 2 [C12O12H2]⋅2H2O (M= Al, Ga, In) MIL-116. Solid synthetic modification of metal–organic frameworks: New types and tandem
State Sci. 26, 38–44. reactions. Coord. Chem. Rev. 378, 500–512.
Wan, J., Shen, Y., Xu, L., Xu, R., Zhang, J., Sun, H., Zhang, C., Yin, C., Wang, X., 2021. Younis, S.A., Bhardwaj, N., Bhardwaj, S.K., Kim, K.-H., Deep, A., 2021. Rare earth
Ferrocene-functionalized Ni (II)-based metal-organic framework as electrochemical metal–organic frameworks (RE-MOFs): Synthesis, properties, and biomedical
sensing interface for ratiometric analysis of Cu2+, Pb2+ and Cd2. J. Electroanal. applications. Coord. Chem. Rev. 429, 213620.
Chem. 895, 115374. Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, M., Lollar, C., Sun, Y., Qin, J., Yang, X.,
Wang, B., Yang, Q., Guo, C., Sun, Y., Xie, L.-H., Li, J.-R., 2017. Stable Zr (IV)-based Zhang, P., 2018. Stable metal–organic frameworks: design, synthesis, and
metal–organic frameworks with predesigned functionalized ligands for highly applications. Adv. Mater. 30, 1704303.
selective detection of Fe (III) ions in water. ACS Appl. Mater. Interfaces 9, Zhang, C., Pan, G., He, Y., 2022b. Conjugated microporous organic polymer as
10286–10295. fluorescent chemosensor for detection of Fe3+ and Fe2+ ions with high selectivity
Wang, H., Chen, Q., Tan, Z., Yin, X., Wang, L., 2012. Electrochemiluminescence of CdTe and sensitivity. Talanta 236, 122872.
quantum dots capped with glutathione and thioglycolic acid and its sensing of Pb2. Zhang, E., Ju, P., Li, Q., Hou, X., Yang, H., Yang, X., Zou, Y., Zhang, Y., 2018a. A novel
Electrochim. Acta 72, 28–31. rhodamine 6G-based fluorescent and colorimetric probe for Bi3+: synthesis,
Wang, L., Luo, D., Qin, D., Shan, D., Lu, X., 2015b. Cathodic electrochemiluminescence selectivity, sensitivity and potential applications. Sens. Actuators B: Chem. 260,
of a CdSe/ZnS QDs-modified glassy carbon electrode and its application in sensing of 204–212.
Pb 2+. Anal. Methods 7, 1395–1400. Zhang, H., Hu, X., Li, T., Zhang, Y., Xu, H., Sun, Y., Gu, X., Gu, C., Luo, J., Gao, B., 2022a.
Wang, L., Zheng, X.-Y., Zhang, X., Zhu, Z.-J., 2021. A quinoline-based fluorescent MIL series of metal organic frameworks (MOFs) as novel adsorbents for heavy metals
chemosensor for palladium ion (Pd2+)-selective detection in aqueous solution. in water: a review. J. Hazard. Mater., 128271
Spectrochim. Acta Part A: Mol. Biomol. Spectrosc. 249, 119283. Zhang, L., Wang, J., Ren, X., Zhang, W., Zhang, T., Liu, X., Du, T., Li, T., Wang, J., 2018b.
Wang, N., Zhao, W., Shen, Z., Sun, S., Dai, H., Ma, H., Lin, M., 2020b. Sensitive and Internally extended growth of core–shell NH 2-MIL-101 (Al)@ ZIF-8 nanoflowers for
selective detection of Pb (II) and Cu (II) using a metal-organic framework/ the simultaneous detection and removal of Cu (ii). J. Mater. Chem. A 6,
polypyrrole nanocomposite functionalized electrode. Sens. Actuators B: Chem. 304, 21029–21038.
127286. Zhang, L., Wang, J., Du, T., Zhang, W., Zhu, W., Yang, C., Yue, T., Sun, J., Li, T., Wang, J.,
Wang, Z., Cohen, S.M., 2009. Postsynthetic modification of metal–organic frameworks. 2019a. NH2-MIL-53 (Al) metal–organic framework as the smart platform for
Chem. Soc. Rev. 38, 1315–1329. simultaneous high-performance detection and removal of Hg2+. Inorg. Chem. 58,
Wang, Z., Babucci, M., Zhang, Y., Wen, Y., Peng, L., Yang, B., Gates, B.C., Yang, D., 12573–12581.
2020a. Dialing in catalytic sites on metal organic framework nodes: MIL-53 (Al) and Zhang, L., Liu, H., Shi, W., Cheng, P., 2019b. Synthesis strategies and potential
MIL-68 (Al) probed with methanol dehydration catalysis. ACS Appl. Mater. applications of metal-organic frameworks for electrode materials for rechargeable
Interfaces 12, 53537–53546. lithium ion batteries. Coord. Chem. Rev. 388, 293–309.
Wang, Z.W., Chen, M., Liu, C.S., Wang, X., Zhao, H., Du, M., Versatile, A., 2015a. AlIII- Zhang, M., Ge, L., Ge, S., Yan, M., Yu, J., Huang, J., Liu, S., 2013. Three-dimensional
based metal–organic framework with high physicochemical stability. Chem. Eur. J. paper-based electrochemiluminescence device for simultaneous detection of Pb2+
21, 17215–17219. and Hg2+ based on potential-control technique. Biosens. Bioelectron. 41, 544–550.
Wilkinson, J., Hooda, P.S., Barker, J., Barton, S., Swinden, J., 2017. Occurrence, fate and Zhang, Y., Yan, B., 2019. A ratiometric fluorescent sensor with dual response of Fe3+/
transformation of emerging contaminants in water: an overarching review of the Cu2+ based on europium post-modified sulfone-metal-organic frameworks and its
field. Environ. Pollut. 231, 954–970. logical application. Talanta 197, 291–298.
Wu, R., Ai, J., Ga, L., 2022. Synthesis of fluorescent copper nanomaterials and detection Zhao, Z.-P., Jiang, Y.-F., Chen, Y., Li, H.-R., Zheng, Y., Zeng, C.-H., Zhong, S., Guo, P.,
of Bi3+. Front. Chem. 10. Zhao, Y.-L., 2018. Highly luminescent lanthanide complex as bifunctional sensor for
Wu, T., Prasetya, N., Li, K., 2020. Recent advances in aluminium-based metal-organic Et2O and Fe2. J. Lumin. 204, 560–567.
frameworks (MOF) and its membrane applications. J. Membr. Sci. 615, 118493. Zhou, H.-C., Long, J.R., Yaghi, O.M., 2012. Introduction to Metal–organic Frameworks.
Xiao-Ping, W., Feng, Y., Fan, D., Jian-Ping, L., 2014. An electrochemiluminescence ACS Publications, pp. 673–674 (in).
biosensor for determination of Pb2+ based on G-quadruplex of aptamer probe. Chin. Zhou, Y., Chen, H.-H., Yan, B., 2014. An Eu 3+ post-functionalized nanosized
J. Anal. Chem. 42, 942–947. metal–organic framework for cation exchange-based Fe 3+-sensing in an aqueous
Xie, D.-h, Ge, X., Qin, W.-x, Zhang, Y.-x, 2021. NH2-MIL-53 (Al) for simultaneous environment. J. Mater. Chem. A 2, 13691–13697.
removal and detection of fluoride anions. Chin. J. Chem. Phys. 34, 227–237. Zhu, Z., He, X., Wang, W.-N., 2020. Unraveling the origin of the “Turn-On” effect of Al-
Xu, L., Gan, S., Zhong, L., Sun, Z., Tang, Y., Han, T., Lin, K., Liao, C., He, D., Ma, Y., 2022. MIL-53-NO 2 during H 2 S detection. CrystEngComm 22, 195–204.
Conductive metal organic framework for ion-selective membrane-free solid-contact
potentiometric Cu2+ sensing. J. Electroanal. Chem. 904, 115923.

25
Another random document with
no related content on Scribd:
WHEN SAILING IN A BOAT, IF NO WIND BLOWS,
YOU FLOAT. AND IF THE BAD SHARK CHASES
AFTER, WHY, GREET HIM WITH SOME HEARTY
LAUGHTER.

1. Having nothing special to do, Uncle Wiggily built himself a sail


boat from a wash tub and some old boards. “Come and take a ride
with me, Nurse Jane,” he invited. “You need a little rest from
washing dishes all the while.” The muskrat lady housekeeper thought
this would be great fun. “But suppose something happens to us,
Uncle Wiggily?” she asked. “It will be an adventure!” laughed the
rabbit.

2. Uncle Wiggily and Nurse Jane took their places in the washtub
boat. “All aboard!” cried Uncle Wiggily. At first the wind blew just
right, and away sailed Uncle Wiggily and Nurse Jane. “A happy
voyage!” grunted Mr. Twistytail, the pig gentleman, waving his
handkerchief at them. “But look out for the Snappy Shark. I heard he
was splashing around in this duck pond ocean!”
3. The wind, blowing strong at first, blew Uncle Wiggily and Nurse
Jane out into the middle of the duck pond ocean. Then, as often
happens, the wind died away and all was calm. “We aren’t moving!”
cried the muskrat lady. “No, we need some wind,” spoke Uncle
Wiggily. “I’ll fan the sail,” offered Nurse Jane. “And I’ll blow on it,”
added Uncle Wiggily. But all this did no good. The ship stood still.
4. While Uncle Wiggily and Nurse Jane were trying to make their
boat go, the Snappy Shark met the Skillery Scallery Alligator near
shore. “Who’s out in that boat?” asked the Shark. “Uncle Wiggily,”
grunted the ’Gator. “Good!” whispered the Shark. “I’ll chase him
from behind, and you swim in front of the boat so he can’t escape
that way. Between us we’ll catch him and nibble his ears. Get ready!”
5. No matter how hard Nurse Jane fanned, nor how hard Uncle
Wiggily blew, they could not, of course, make wind enough to move
the boat. Then the bunny said: “I’ll do as the sailors do. I’ll whistle a
tune and see if that will bring a breeze!” So Uncle Wiggily began to
whistle, but Nurse Jane cried: “Stop! You have called the Snappy
Shark instead of the wind! Oh, what shall we do to get away?”
6. “Don’t be afraid!” cried Uncle Wiggily, when Nurse Jane told
about the Shark. “If the wind won’t blow our boat I’ll jump in and
swim ashore, pulling the boat after me with the anchor and rope.”
Just as the bunny was about to jump in, he saw, in front of his ship,
the Skillery Scallery Alligator swimming along. “Ah, I have a better
plan!” laughed the bunny. “I’ll make the ’Gator pull us to shore!”
7. Uncle Wiggily dropped the anchor and tow rope so that it caught
on one of the ’Gator’s legs, and then the Skillery Scallery creature had
to pull the boat along, whether he wanted to or not. “Stop! Stop!
Stop!” cried the Shark. “How can I nibble ears when you are pulling
Uncle Wiggily away?” But the ’Gator would not stop, not knowing
any better. “I’ll cut the rope when we are near shore!” said the bunny.
8. “What are you doing, Nurse Jane?” asked Uncle Wiggily, as the
’Gator towed the boat faster and faster. “Trying to sprinkle pepper in
the eyes of the Shark,” answered the muskrat lady. “Never mind
about that,” said Uncle Wiggily. “We are close to the dock now, and
I’ll cut the rope so we’ll drift in.” Then the bunny did this, and then
he had to fan Nurse Jane, who fainted. Now what comes next?
9. Safely to the dock floated Uncle Wiggily’s boat, which had been
towed by the ’Gator, even if there was no wind. “Oh, such a voyage!”
cried Nurse Jane, as Mr. Twistytail helped her off the ship. “But see
what happened to the Shark and ’Gator!” laughed Uncle Wiggily.
“Their plans went wrong. They bunked together and how angry they
are!” The Shark blamed the ’Gator and the ’Gator blamed the Shark!
And if the lead pencil doesn’t give the fountain pen a black mark for
staying out too late at the moving pictures, you shall next hear about
Uncle Wiggily’s Bob Sled.
UNCLE WIGGILY MADE A BOB SLED TO GIVE HIS
FRIENDS A SLIDE; BUT QUICKLY DUMPED THE
BUSHY BEAR WHO TRIED TO STEAL A RIDE.

1. One day, early in the new year, Nurse Jane Fuzzy Wuzzy heard
queer, pounding noises out in the wood shed. “I hope that isn’t the
Fuzzy Fox trying to break in to nibble ears,” she said. And when she
looked in the shed she saw Uncle Wiggily and Uncle Butter making a
coasting bob sled. “You are too old for such fun!” laughed Nurse
Jane. “I’ll show you how to slide down hill!” said Uncle Wiggily.
2. At last the bob sled was finished and Uncle Wiggily and his goat
friend started pulling it toward the coasting hill. On the way they saw
Aunt Lettie. “Come on,” spoke Uncle Wiggily, “we’ll give her a ride.
She’ll tell Nurse Jane about it and my muskrat lady housekeeper will
know I’m a good sled-maker.” Up behind Aunt Lettie the two friends
pulled the bob. “How you startled me!” she bleated. “Baa-a-a!”
3. “How are you enjoying yourself, Aunt Lettie?” asked Uncle
Wiggily, as the goat lady sat on the bob and was hauled away. “Oh, I
am having a lovely time, thank you,” she bleated. “But there is poor
Mrs. Twistytail just ahead. She is so fat she has to sit down to rest her
feet.” Uncle Wiggily twinkled his pink nose and said: “Why not give
the lady pig a ride? We have plenty of room.” So they asked her.
4. Mrs. Twistytail was very glad to get on the bob sled of Uncle
Wiggily, and while the two animal ladies were enjoying the ride, the
Bob Cat and Bushy Bear happened to see them. “Oh, I know how we
can catch Uncle Wiggily!” whispered the Bear. “How?” asked the Bob
Cat. “I’ll tell you in my den,” growled the Bear. “It’s time we had
some nibbles off that rabbit’s ears. This time we’ll get him!”
5. To the Bear’s den the two bad chaps hurried. “Come, Wife!”
growled the bushy bruin fellow, “give me some of your old dresses, a
bonnet and a shawl!” Mrs. Bruin wanted to know if they were to be
sold to the rag man. “No!” growled Mr. Bruin. “I’m going to dress up
like an old lady. Uncle Wiggily will invite me to ride on his bob sled.
When I’m on I’ll jump off and nibble his ears. Give me a dress!”
6. When the Bushy Bear was dressed in some of his wife’s old
clothes, he went out and stood in the road, turning his back. “You
look just like a poor old woman!” whispered the hidden Bob Cat. And
when the bunny and Uncle Butter came along with the bob sled, the
rabbit said: “Look, there’s another poor, old, tired lady. We’ll give her
a ride. We must be kind to the old folks. Hop on, lady!” he cried.
7. Keeping the bonnet pulled down over his face, the Bushy Bear,
pretending to be a lady, got on the bob sled. “Put your paws around
me and hold on,” invited Mrs. Twistytail. But as soon as the Bear did
that, he showed his long claws, which the pig lady saw. “Oh, mercy!”
squealed Mrs. Twistytail. “This is dreadful!” The bunny and the goat,
who were giving the ladies a ride, turned to see what was the matter.
8. “This isn’t a poor old lady at all!” squeaked Mrs. Twistytail. “It’s
the Bushy Bear with a dress on!” Then Uncle Wiggily saw that he had
been fooled. “Quick, Uncle Butter!” whispered the rabbit. “Give the
rope a hard pull! We’ll get rid of the bad chap!” Jerking on the sled
suddenly, the Bear was jiggled off backward. “Now let’s run to the
hill and coast down!” cried the bunny. “Run fast, Butter!”
9. Away ran the bunny and the goat, pulling the sled with the
animal ladies on it. Casting aside his wife’s dress, which tangled in
his legs, the Bushy Bear tried to follow. But Uncle Wiggily soon
reached the top of the coasting hill. “Down we go!” he cried as he and
the goat jumped on the sled and slid to safety. “Fooled again!”
growled the Bear, sitting down on top of the hill. “Plop him!” cried
the Squigglers.
When you have finished reading this nice little book, perhaps you
would like to read a larger volume about Uncle Wiggily.
If so, go to the book store and ask the Man for one of the Uncle
Wiggily Bedtime Story Books, they have a lot of Funny Pictures in
and 31 stories—one for every night in the month. If the book store
man has none of these volumes ask him to get you one or send direct
to the Publishers,

A. L. BURT COMPANY,
114 EAST 23rd STREET
NEW YORK CITY

You might also like