Open access peer-reviewed chapter - ONLINE FIRST

Carbazole-Based Schiff Bases: Structural Insights and Applications toward Metal Ion Detection

Written By

Syeda Aaliya Shehzadi and Mustaghees Ur Rehman

Submitted: 17 May 2025 Reviewed: 21 May 2025 Published: 09 July 2025

DOI: 10.5772/intechopen.1011144

Schiff Bases - Recent Developments and Application Areas IntechOpen
Schiff Bases - Recent Developments and Application Areas Edited by Nuriye Tuna Subasi

From the Edited Volume

Schiff Bases - Recent Developments and Application Areas [Working Title]

Assistant Prof. Nuriye Tuna Subasi

Chapter metrics overview

3 Chapter Downloads

View Full Metrics

Abstract

Carbazole-based Schiff bases have emerged as versatile chemosensors for detecting toxic and essential metal ions, addressing critical environmental and health concerns arising from metal pollution. This chapter explores their structural design, emphasizing the integration of the electron-rich carbazole core with imine (‒C=N‒) functionalities, which enhance π-conjugation, thermal stability, and selective metal coordination. Detection of transition metal ions such as Fe3+, Cr3+, Al3+, and Cu2+ has been successfully achieved using carbazole-based Schiff bases with a detection limit up to nanomolar range. Heavy metals, such as Hg and Pb, have also been detected using such materials. These sensors operate via mechanisms such as photoinduced electron transfer (PET), excited-state intramolecular proton transfer (ESIPT), chelation-enhanced fluorescence (CHEF), and aggregation-induced emission enhancement (AIEE), enabling “turn-on” or “turn-off” optical responses. This chapter highlights advancements in logic gate integration, smartphone-based detection, and AIEE-active probes for on-site monitoring. Challenges such as hydrolytic instability and aqueous solubility are addressed through structural modifications, including sulfonate groups or nanomaterial integration. Future directions emphasize push-pull architectures, near-infrared emission, and user-friendly formats like hydrogel strips. By bridging synthetic versatility with functional adaptability, carbazole-Schiff bases offer scalable solutions for environmental, industrial, and biomedical metal sensing, underscoring their potential in next-generation detection technologies.

Keywords

  • chemosensor
  • carbazole
  • Schiff base
  • metal ion detection
  • fluorescent probes

1. Introduction

The increasing industrialization and anthropogenic activities have led to the widespread distribution of heavy and transition metal ions in various ecosystems, posing serious threats to both environmental and human health. Thus, detection of metal ions in environmental and biological systems has emerged as a critical area of research, driven by the dual imperatives of safeguarding human health and preserving ecological balance.

Heavy metals such as mercury (Hg2+), lead (Pb2+), and chromium (Cr3+/Cr6+) are notorious for their toxicity even at trace concentrations, contaminating water supplies, soil, and food chains. Conversely, essential ions like iron (Fe3+), zinc (Zn2+), copper (Cu2+), and cobalt (Co2+) play vital roles in biological processes but become harmful when dysregulated. Traditional analytical methods for metal ion detection, including atomic absorption spectroscopy (AAS) and inductively coupled plasma mass spectrometry (ICP-MS), offer high sensitivity but suffer from limitations such as costly instrumentation, complex sample preparation, and lack of real-time monitoring capabilities. This has spurred interest in developing chemosensors—molecular probes that combine selectivity, sensitivity, and practicality for on-site detection. Among the plethora of sensing platforms, Schiff bases, being dynamic ligands formed via condensation of amines (NH2) and carbonyl (C=O) groups, have gained prominence due to their synthetic versatility, strong metal-binding affinities, and tunable optical responses. Schiff bases are typically formed via condensation reactions between primary amines and carbonyl compounds, resulting in imine (‒C=N‒) functionalities that serve as key binding sites for metal ions.

In recent years, Schiff base sensors and their metal ion complexes have gathered significant attention as a prominent research focus, driven by their diverse practical applications across scientific disciplines. These include environmental monitoring and biological cell imaging, where they enable precise detection of target ions within cellular structures, tissues, and organelles [1, 2, 3, 4, 5]. Additionally, their customizable design and multi-stimuli-responsive characteristics make them valuable in optoelectronic systems [6, 7, 8]. The inherent molecular switching behavior of Schiff bases has further been exploited in advanced technologies such as molecular keypads and logic gates [9, 10].

The carbazole moiety, a nitrogen-containing heterocycle with an electron-rich aromatic core, provides a rigid, planar scaffold that enhances π-conjugation, thermal stability, and luminescent properties. When integrated with the imine (‒C=N‒) group, a hallmark of Schiff bases, these hybrid systems exhibit synergistic effects: the carbazole unit amplifies fluorescence or colorimetric signals, while the imine group acts as a selective coordination site for metal ions. This dual functionality enables carbazole-Schiff bases to act as “turn-on” or “turn-off” sensors, with detectable changes in absorption or emission spectra upon metal binding. Moreover, their modular synthesis allows for precise structural modifications, such as introducing electron-donating/withdrawing substituents or extending conjugation, to optimize sensitivity toward specific ions.

Over the past decade, researchers have harnessed these attributes to design carbazole-Schiff base sensors for detecting a wide array of metal ions. For instance, derivatives functionalized with hydroxyl (-OH) or thiol (-SH) groups demonstrate a high affinity for Hg2+, a potent neurotoxin, via soft-soft interactions with sulfur or nitrogen donors. Similarly, carboxylate-appended (-COO¯) carbazole-Schiff bases selectively bind Fe3+ through chelation-enhanced quenching (CHEQ) mechanisms, enabling quantification in biological fluids [11]. The selectivity of these sensors often arises from the geometric and electronic complementarity between the ligand’s binding pocket and the target ion. Beyond selectivity, the sensitivity of carbazole-Schiff bases, often achieving detection limits in the nanomolar range, is attributed to their amplified signal transduction. The extended π-system of carbazole facilitates intramolecular charge transfer (ICT) or photoinduced electron transfer (PET) processes, which are perturbed upon metal coordination, leading to pronounced optical changes.

Recent innovations, such as incorporating aggregation-induced emission (AIE) motifs or graphene oxide nanocomposites, have further improved detection thresholds, enabling trace-level analysis in complex matrices like wastewater or cellular environments [12, 13]. Although some extensive review articles have addressed the potential of Schiff bases and their derivatives toward metal ion detections, [14, 15, 16, 17, 18] a comprehensive summary specifically addressing the potential of carbazole-based Schiff bases and their use in this context remains absent in the literature.

The current chapter provides a comprehensive exploration of carbazole-based Schiff bases as next-generation metal ion sensors. It begins by elucidating the structural design principles that govern their sensing behavior, emphasizing the interplay between the carbazole core, Schiff base linker, and auxiliary functional groups, with detailed discussion on the mechanisms of ion recognition, mechanisms underlying metal ion recognition, spectroscopic responses including chelation dynamics, electronic interactions, and steric effects, followed by structure-activity relationships.

Through comprehensive literature examples, the emerging trends and opportunities in this rapidly evolving field are identified and highlighted. By reviewing recent literature and critically analyzing structure-property relationships, this chapter aims to inspire innovative approaches to sensor design while bridging the gap between academic research and industrial implementation.

Advertisement

2. Detection of transition metal ions (Fe, Cr, Al, Cu)

Transition metal ions are among the most widely invested metals worldwide. Many industries, such as mining, electronics, and battery manufacturing, release metal ions as waste. They can contaminate water, soil, and food chains if not treated and regulated, causing serious diseases in humans and animals. Strict regulatory limits on metal concentrations by government bodies demand sensitive detection methods, essential for compliance and advancements in sensors, catalysis, and materials science.

2.1 Detection of Fe3+ and Cr3+

Fe3+ is vital for biological processes like oxygen transport and cellular metabolism, but its imbalance (deficiency or excess) leads to disorders such as anemia, organ damage, and diabetes, highlighting the need for precise detection methods to monitor its levels for human health. Li et al. [19] have synthesized a carbazole-based Schiff base fluorescent probe by reacting f 3-(2-formyl)thienyl-9H-hexylcarbazole (1) with 4-hydroxy-3-nitroaniline (2a) to give probe 1a (P1a) and 3-nitroaniline (2b) to give probe 1b (P1b). The thiophene group was introduced into the 3-position of carbazole to extend the conjugated π-bridge and also to provide a new coordination binding site from the sulfur atom of thiophene, as shown in Figure 1. The conjugation extended from carbazole moiety to thiophene to benzene ring via ‒C=N‒ group. The structures of synthesized probes were confirmed via IR and NMR spectroscopies. Probe 1a showed dramatic enhancement of the fluorescent intensity upon addition of Fe3+ or Cr3+ in the presence of other metal ions, such as Na+, Ni2+, K+, Ag+, Ca2+, Cd2+, Hg2+, Zn2+, Co2+, Mn2+, Mg2+, and Fe2+, which had no distinct influence on fluoresce intensity. The emission peak was red-shifted from 495 to 502 nm. While probe 1b showed an increase in fluorescent intensity in almost all other ions, suggesting that 1b had no selective recognition for the tested metal ions. This suggested that chemosensors without a phenolic –OH group in the ortho position exhibited weak selective recognition for tested metal ions, indicating that the O atom of –OH is the significant recognition site for metal ions.

Figure 1.

The synthetic route of probe 1 (P1a-b) and binding mode of P1a with M (M = Fe3+ and Cr3+).

The fluorescence enhancement of P1a upon binding with Fe3+ or Cr3+ was perhaps due to inhibition of -C=N isomerization and ESIPT processes that typically quench fluorescence in the unbound state; the selective coordination with these metal ions stabilized the molecular structure, suppressing non-radiative decay pathways and resulting in increased fluorescence intensity. The authors also investigated the binding stoichiometry and binding mode of P1a for both Fe3+ and Cr3+. The binding stoichiometry was 2:1 for both ions (P1/M3+) as determined by Job’s plot. Binding mode was confirmed by 1H-NMR, upon coordination with Cr3+, the proton signals of P1a broadened and exhibited downfield shift for the hydroxyl, imino, and thienyl protons, indicating that the oxygen atom of the phenolic –OH, the nitrogen atom of the imine group (–N=CH), and the sulfur atom of the thienyl ring served as key binding sites, as presented in Figure 1. The absence of the ortho –OH group in P1b correlates with its diminished metal ion recognition ability.

By increasing the number of nitrogen coordination sites, a much stronger effect can be expected. He et al. [20] conceived this idea and synthesized a dimeric Schiff-base probe 2 (P2) by reacting 9-hexylcarbazole-3-carbaldehyde (3) with 2,3-diaminomaleonitrile in ethanol under reflux conditions as presented in Figure 2. The hexyl chains enhanced the solubility in organic solvents like DMF, while the Schiff base (‒CH=N‒) and nitrile (‒CN) groups provided coordination sites for metal ions. Both absorbance and fluorescence spectra were studied to find the selective response of probe P2 to a variety of 18 metal ions.

Figure 2.

Synthesis and binding mode of probe P2 with Fe3+.

P2 exhibited high selectivity for Fe3+ over 17 other metal ions, as shown in Figure 3, during fluorescence studies in DMF with 60% quenching at 372 nm exclusively with Fe3+, with no significant response to competing ions. Competitive experiments confirmed minimal interference. The detection limit obtained was 3.75 × 10−8 M, which is well below the EPA’s permissible limit (5.35 μM) for drinking water. Job’s plot and Benesi-Hildebrand analyses established a 1:1 binding stoichiometry between P2 and Fe3+. The association constant (Ka = 7.98 × 106 M−1) indicated strong affinity. Binding occurred via coordination of Fe3+ to both nitrogen atoms of the imino group (‒C=N‒), forming a rigid 5-membered chelate (Figure 2). This interaction disrupted the ligand’s excited-state processes, leading to fluorescence quenching. In the unbound state, P2 emitted strongly due to restricted C=N isomerization and ICT. Fe3+ coordination induced ligand-to-metal charge transfer (LMCT) by dissipating energy non-radiatively and thus quenching fluorescence. Reversibility with EDTA confirmed that the process is dynamic and practical for reuse.

Figure 3.

UV response of P2 with Fe3+ and other metal ions in DMF solvent, adapted from Ref. [20].

Iron detection by imino nitrogen and phenolic oxygen is highly pragmatic, as in another report by Nandhakumar et al. [21], a carbazole-derived Schiff base fluorescent chemosensor (P3) was synthesized via condensation of 9-ethyl-9H-carbazol-3-amine (4) and salicylaldehyde (5) in ethanol, yielding a 73% of P3 as summarized in Figure 4. The structure was characterized by NMR, IR, and mass spectrometry.

Figure 4.

Synthesis of P3 and binding mechanism toward Fe3+.

The authors tested P3 for the detection of Fe3+ and arginine amino acids. The synthesized probe exhibited a sequential “on-off-on” fluorescence response, where Fe3+ induced a turn-off via ICT from the carbazole nitrogen to Fe3+, while arginine (Arg) restored fluorescence (turn-on). Job’s plot and Benesi-Hildebrand analysis revealed a 2:1 (P3:Fe3+) binding stoichiometry with a binding constant Ka = 2.09 × 104 M−1 and a low detection limit (LOD) of 12.22 nM for Fe3+. Coordination involved the carbazole nitrogen, imine nitrogen, and phenolic oxygen, confirmed by FT-IR and 1H-NMR shifts. The sensor operated in DMF/H2O (1,1, pH 7.4) with reversibility via EDTA and pH stability (6–8). Authors also investigated the detection ability of P3 in molecular logic gates smartphone-based on-site detection of Fe3+ & Arg via RGB colorimetry. The detection of Fe3+ ion in A549 cells using live-cell imaging showed quenching of fluorescence upon treating cells with FeCl3 solution, while a real-sample Fe3+ analysis in water with 99–101% recovery was achieved.

2.2 Detection of Al3+

Aluminum is the third most abundant element in the Earth’s crust, comprising roughly 8% by weight, and is extensively used across industries such as transportation (aircraft, cars), construction (window frames, roofing), packaging (beverage cans, foil), and electrical applications (power lines, wiring) and in many consumer goods like cookware, appliances, and furniture. Exposure to aluminum is usually not harmful, but excessive contact with Al3+ can bioaccumulate in the body and is implicated in neurological disorders (e.g., Alzheimer’s and Parkinson’s disease), kidney damage, chronic renal failure, and bone softening due to its interference with physiological processes [22, 23]. Consequently, developing highly selective and sensitive Al3+ detection methods is essential for monitoring environmental and biological systems and mitigating their adverse health and ecological impacts. A challenge in developing chemosensors for Al3+ is its lower coordination ability than other transition metals, and it is difficult to identify spectroscopically due to its lack of photophysical characteristics. However, O- and N- being strong hetero donor centers could capture Al3+ in soluble form.

Another design similar to P3 was reported by Babu et al. [24] based on Schiff base, derived from 4 and 3,5-diiodo salicylaldehyde (6) in ethanol, as shown in Figure 5. The strong electron donor effect of two iodo groups in P4 could enhance the donor ability of the =OH group.

Figure 5.

Synthetic route to probe 4 (P4).

In 90% aqueous CH3CN, the probe P4 exhibited a weak emission at 540 nm (λex = 385 nm) due to non-radiative decay via ESIPT and -C=N isomerization. Among various transition metal cations added, such as Sn2+, Cd2+, Mn2+, Al3+, Cr3+, Co2+, Cu2+, Hg2+, Pb2+, Zn2+, and Fe3+, only Al3+ induced a ~ 6-fold “turn-on” fluorescence with a red shift to 559 nm and the appearance of new absorption bands at 428 and 496 nm, which was attributed to ground-state complex formation and Metal to Ligand Charge Transfer (MLCT) transitions. Benesi-Hildebrand and Job’s plot analyses revealed a 1:1 of P4-Al3+ stoichiometry with Ka = 5.44 × 104 L mol−1. The LOD was determined as 0.403 mg μL−1. The mechanism of binding involved the coordination of phenolic O- and imine N-donors, conferring high selectivity for Al3+ over a panel of competing cations. The authors did not propose a binding mode or bounded structure of complex.

Replacing diiodo salicylaldehyde with naphthaldehyde led to enhanced fluorescence due to extended π-conjugation, as witnessed by a carbazole Schiff-based Al3+ detector, reported by Kaya et al. [25] where they condensed 4 and 2-hydroxy-1-naphthaldehyde (7) in ethanol under refluxed conditions to get probe P5 as shown in Figure 6. They tested the probe P5 for both cations (K+, Ag+, Ba2+, Mn2+, Mg2+, Sn2+, Hg2+, Ca2+, Co2+, Zn2+, Cu2+, Ni2+, Pb2+, Al3+, Fe3+, Cr3+, and Cr6+) and anions (F, Cl, Br, I, CO32−, HCO3, and HPO4) in aqueous media. The probe demonstrated high selectivity toward Al3+ over 16 competing metal cations and seven anions.

Figure 6.

Synthesis and binding mode of probe 5 (P5) with Al3+.

Fluorescence studies revealed an 11-fold emission enhancement at 533 nm upon Al3+ addition, with no significant response to other ions. Competitive experiments confirmed that Al3+ detection remains unaffected in the presence of interfering species, revealing its specificity. The binding stoichiometry between P5 and Al3+ was recognized as 1:1 using Job’s plot and Benesi-Hildebrand analyses. The Ka value was calculated as 5 × 104 M−1, indicating strong affinity. Binding occurred via coordination of Al3+ to the imine nitrogen (‒C=N‒) and the deprotonated hydroxyl oxygen (O-), forming a rigid, planar complex as presented in Figure 5. This interaction suppresses excited-state intramolecular proton transfer (ESIPT) and C=N isomerization, which is a key non-radiative decay pathway in the free ligand. The mechanism of detection was the “turn-on” fluorescence response based on chelation-enhanced fluorescence (CHEF) effect. Authors believed that a bathochromic shift in UV-Vis spectra from 462 to 479 nm on binding showed the involvement n → π* electronic transition due to lone pairs of nitrogen and oxygen atoms resulting in a six-membered ring after chelation. The LOD calculated was 2.59 × 10−7 M, below the WHO-recommended threshold for Al3+ in drinking water (2.41 μM), suggesting its utility in practical applications.

In another report by Kaya et al. [26] the selective Al3+ detection was achieved by synthesizing carbazole Schiff-base probe 6 (P6) having benzene ring as π-linker between carbazole and imino group while introducing ‒OH and ‒OCH3 groups on aldehyde partner. The synthesis pathway is shown in Figure 7. The electron-donating ability of methoxy group enhanced the coordination power of the designed probe for low-coordinating metals. The P6 was tested for various metal ions (Co2+, Cu2+, Ni2+, Fe3+, Cr3+, K+, Zn2+, Mn2+, Al3+ and Ag+) by fluorescence study, it exhibited weak fluorescence intensity (λex = 320 nm, λem = 465 nm) in free form due to ‒C=N isomerization and ESIPT, but the fluorescence intensity increased 108 times with Al3+, with no prominent change for other metal ions.

Figure 7.

Synthetic pathway and binding mechanism of P6 for Al3+.

The LOD was 9.29 × 10−7 M, and Ka value using the modified Benesi-Hildebrand equation was 1.64 × 104 M−1. The binding stoichiometry of the complex between P6 and Al3+ was stated at a ratio of 2:1, but in binding mechanism, it showed 1:1. Although this sensor worked well in DMSO/H2O (1:1) environment, most of such chemosensors suffer from water solubility issues. This limitation makes them impractical for real-world problems. Water is a strong solvent, and its interactions with Al3+ (through hydration) can outcompete the interactions with potential ligands in a detector. This usually results in weak, unstable, or even insoluble Al3+ complexes, making it difficult to create a reliable detector. Thus, the development of a stable, water-soluble Al3+ ion detector remains a challenge due to its lower coordination tendency and leading to weak interactions in water. Chemli et al. [27] has devised a symmetric optical probe (P7) by condensing 2,7-bisformyl-N-pentylcarbazole (8) and 2-aminophenol (9) moieties in methanol at 80°C for 4 h to give P7 as shown in Figure 8. Structure was confirmed by IR, Mass, and NMR spectroscopic methods.

Figure 8.

Synthetic route toward probe 7 (P7).

This design led to AIEE utilizing ESIPT in THF/H2O mixtures, with enhanced fluorescence observed at 480 nm when the water fraction reached 70%. Upon addition of Al3+, a distinct color change from yellow to colorless was observed under daylight, and an intense blue emission under UV light, indicating a highly selective “turn-on” fluorescence response. Among a panel of metal ions, only Al3+ caused significant spectral changes. Fluorescence and UV-Vis titrations confirmed a 1:1 binding stoichiometry (Job’s plot) and an association constant (Ka) of 1.15 × 104 M−1 (Benesi-Hildebrand method). The LOD was determined as 112 nM (UV-Vis) and 9.4 nM (fluorescence), well below the WHO guideline of 7.4 μM for Al3+ in drinking water.

DFT calculations revealed that the regions with high electronegativity, having the strongest attraction to cations binding sites, were observed around the nitrogen atom of the imine groups (‒C=N‒) and the oxygen of the phenol groups (‒OH) (Figure 9). The binding mechanism was attributed to Al3+-induced hydrolysis of the imine bond, regenerating 2,7-bisformylcarbazole, as confirmed by 1H-NMR spectral shifts and fluorescence quantum yield, which increased from 0.5% (Carb-Azo) to 13.7% (product). Furthermore, a smartphone-assisted fluorometric platform using a fiber-optic endoscope was developed for practical Al3+ detection in drinking water, achieving a detection limit of 0.4 μM. Recovery experiments in spiked commercial water samples gave values from 94.8 to 116.7% with RSDs <12%, validating the practical applicability of both the spectroscopic method and the portable device for real-time environmental monitoring of Al3+ contamination.

Figure 9.

Binding mode of P7 and Al3+ with 2:1 stoichiometry of P7:Al3+.

Inclusion of benzophenone moiety on carbazole not only can extend conjugation but also enhance the photophysical properties of probe overall, owing to its strong electron-accepting and UV-absorbing abilities. Hu and Tao group [28] recently designed an AIE-active Schiff-base probe, P8, tailored for selective aqueous Al3+ detection. The P8 was synthesized by condensing a carbazolyl-benzophenone intermediate (8) with 5 in refluxing toluene. The structure was confirmed by 1H/13C NMR showing imine (‒CH=N‒) CH signal at ~8.7 ppm δ-value and phenolic OH at ~12.9 ppm δ-value (Figure 10).

Figure 10.

Synthesis pathway toward probe P8.

P8 exhibited classic aggregation-induced emission in THF/H2O, 1:4. As the water fraction increased from 0 to 95%, its UV–Vis π–π* band red-shifted from 380 to 410 nm and its emission at 545 nm intensified 6.4-fold, driven by aggregation-caused restriction of intramolecular rotations and hydrogen-bond–mediated planarization. Upon addition of Al3+, a pronounced “turn-on” fluorescence at 545 nm and a new absorption at 410 nm emerged, while 18 competing metal ions (Li+, Na+, K+, Ag+, Cu2+, Fe3+, Zn2+, Mn2+, Ca2+, Mg2+, Cd2+, Cr2+, Pb2+, Ba2+, Ga2+, La3+, Ni2+, In3+, Ce3+) elicited negligible response, owing to chelation-enhanced fluorescence (CHEF) that inhibited C=N isomerization and ESIPT quenching by coordinating Al3+ to the imine nitrogen and phenolic oxygen donors. FT-IR confirmed this binding by the disappearance of the 3051 cm−1 OH stretch and shift of the 1600 cm−1 C=N band to 1586 cm−1, and 1H-NMR titrations revealed loss of the OH singlet and downfield imine CH, verifying O/N chelation. Job’s plot analysis indicated a 1:1 P8-Al3+ stoichiometry, and Benesi-Hildebrand fitting gave Ka = 0.477 × 103 M−1 (Figure 11).

Figure 11.

Binding mechanism of P8 with Al3+.

Fluorescence titrations established a linear detection range of 50–500 μM and a low detection limit of 1.486 μM, which is well below the WHO guideline of 7.41 μM. Application to real water samples showed recoveries of 100.72–102.85% with RSD <2.82%, revealing P8’s potential for sensitive, selective, and practical environmental monitoring of Al3+.

2.3 Detection of Cu2+

Copper is the third most abundant transition metal in the human body and is vital for the development and function of multiple organs such as the heart, central nervous system, immune system, and liver, yet its dysregulation can lead to disorders like Menkes and Wilson’s diseases when deficient or overloaded [29]. Alongside its biological importance, copper has wide utilization in different industries like conducting wires, batteries, switches, transformers, and telecommunications equipment. Therefore, developing highly sensitive and selective chemosensors for trace Cu2+ detection is critical. Recent advances in fluorescent probes demonstrating “turn-on” mechanisms and detection limits in the micromolar to sub-micromolar range envisaged the idea of π-bridging between molecules to make their absorption and emission bands red-shifted, and then the color variation can be observed and determined via naked eyes [30, 31]. In view of this concept, Zhu and Fang et al. have disclosed [32] a report where they introduced the 4-aminobenzyl cyanide unit into the 3-position of carbazole to extend the π-conjugated bridge (9), the –NH2 handle made Schiff base and water solubility boosted via cyano group. The N,N-diethyl-bearing salicylaldehyde (10) was then condensed to yield the Schiff-base probe 9 (P9), enabling metal coordination through the imine (‒C=N‒) nitrogen and phenolic oxygen. The whole design and synthesis are shown in Figure 12. The sensing ability of the final probe was investigated in CH3CN-Tris (20 mM, v/v = 1:1, pH = 7.2) solution for a range of 13 metal ions (Na+, K+, Ca2+, Mg2+, Ni2+, Fe3+, Co2+, Cu2+, Zn2+, Hg2+, Ag+, Cd2+, Pb2+, and Al3+) via absorption spectra. The UV-Vis spectrum of P9 displayed π–π* and ICT absorption bands at 288 and 430 nm, respectively. On Cu2+ addition, the main band blue-shifted to 418 nm with markedly decreased intensity and a color change from bright yellow to near-colorless, whereas other metal ions induced no or minor spectral shifts (Figure 13).

Figure 12.

Synthetic route of P9.

Figure 13.

Sensing mechanism and binding of P9 with Cu2+.

In fluorescence spectroscopy, a bright-yellow fluorescence (λex = 420 nm, λem = 520 nm) was observed, which was selectively and dramatically quenched by Cu2+ ion, accompanied by a visible color change from yellow-green to colorless, while other metal ions induced negligible spectral changes. UV-Vis titrations and Benesi-Hildebrand analysis revealed an association constant Ka = 1.47 × 105 M−1 and a 1:1 P9-Cu2+ stoichiometry (confirmed by Job’s plot), with a fluorescence-based detection limit of 1.14 × 10−5 M; the rapid (≤30 s) and reversible (via EDTA) response. 1H-NMR titrations reveal downfield shifts of imine and phenolic protons, pinpointing coordination through the nitrogen of ‒C=N and phenolic oxygen, while ESI-MS and DFT calculations supported the formation of a non-fluorescent ground-state complex that opens a photoinduced electron transfer (PET) quenching pathway. P9 was also deployed on test strips for naked-eye on-site Cu2+ detection in aqueous media and demonstrated cellular permeability via Cu2+-responsive “turn-off” imaging in HeLa cells, highlighting its dual colorimetric and fluorometric sensing capabilities for environmental and biological applications.

Gosh et al. [33] envisioned another idea to find selectivity for Cu2+ by the use of Low Molecular Weight Gelators (LMWGs) as sensory material. For that purpose, they reacted the naphthalimide aldehyde (11) with 4 under reflux in ethanol to give probe 10 (P10) an 85% yield. The P10 formed thermo-irreversible gels with minimum gelation concentration (5 mg mL−1 in DMF/H2O, DMSO/H2O, and 1,4-dioxane/H2O (1:1 v/v)). Gel-sol transition temperatures (T9el) were 61, 60, and 56°C, respectively. Morphology and rheology revealed fibrous networks whose packing varied with solvent.

Among the different metal ions (Hg2+, Ag+, Fe2+, Cu2+, Fe3+, Pb2+, Cd2+, Al3+, Ca2+, Ni2+, and Co2+) tested, the gel underwent a gel-to-sol transition upon addition of ≥0.5 equiv. Cu2+ or Fe3+ in DMF/H2O as solvent. Subsequent treatment with F distinguished Fe3+ (greenish-yellow) from Cu2+. The solution phase sensing showed a π–π* absorption at 455 nm and green emission at 520 nm; both were selectively quenched by Cu2+ and Fe3+ (other ions inert). Job’s plots indicated a 1:2 (M/P10) stoichiometry, and emission titrations yielded binding constants for Cu2+ and Fe3+ as Ka = (4.6–6.9) × 103 M−1, Ka = (5.3–5.8) × 104 M−1 respectively. Detection limits were 2.02 × 10−7 M (Cu2+) and 2.09 × 10−7 M (Fe3+). Single-crystal X-ray of the Cu2+-complex revealed a square-planar coordination via two imines (‒C=N‒) N and two phenolate O donors, as shown in Figure 14. DFT/TD-DFT reproduced the disappearance of the 455 nm band upon complexation and mapped HOMO-LUMO transitions consistent with ICT-driven quenching. The extended conjugation and polar imide groups of this design enhanced the fluorescence ability and water solubility, a requirement for practical and industrial applications.

Figure 14.

Synthetic route to P10 and its binding mechanism to Cu2+.

The conjugation can be extended via polyaromatic compounds like naphthalene and incorporating heterocyclic compounds as condensing partners to increase the chances for colorimetric recognition and improved solubility. Zhu et al. [34] synthesized probe 11 (P11) by Suzuki coupling of 3-formyl-6-iodo-N-butylcarbazole (12) with 2-naphthylboronic acid (13) to give 14, which upon condensation with 2-hydrazinylbenzothiazole (15) in ethanol yielded P11 as shown in Figure 15. Structural characterization via1H-NMR, FT-IR, and ESI-MS confirmed the imine (-CH=N-) linkage and naphthalene extension, enhancing π-conjugation. They tested the sensing ability of designed probe P11 for variety of alkali, alkaline-earth, transition, and heavy metal ions. The compound selectively detected Co2+ colorimetrically in DMSO/H2O (1:1 v/v) via a visible color shift (colorless → pale yellow) with a new absorption band at 436 nm, attributed to LMCT. For Cu2+, it functioned as a fluorescent “turn-on” sensor in CH3CN, showing a 10-fold emission enhancement at 418 nm with a blue shift, driven by suppression of PET and activation of CHEF. The chemical binding ratio of CNS to Co2+/Cu2+ was 1:1 according to Job’s plot, while the LOD for Co2+/Cu2+ was 4.31×10−7 and 1.98×10−6 molL−1, respectively. 1H-NMR and DFT studies identified binding sites at the imine nitrogen and thiazole ring nitrogen, with Cu2+ coordination stabilizing the complex via LMCT, as shown in Figure 16.

Figure 15.

Synthetic route to P11.

Figure 16.

Binding mechanism of P11 with Co2+ and Cu2+.

In another report by Mishra et al. [35] a turn off Cu2+ sensor based on bis-carbazole and bis-anthracene Schiff base was synthesized by reacting 9-hexylcarbazole-3-carbaldehyde (16) and ethane-1,2-diamine, leading to two imino C=N‒ functionalities in the probe as presented in Figure 17. Due to relevance, only carbazole-based probe (P12) is discussed here. The P12 was tested for nitrate salts of 12 different metals (Fe2+, Cr3+, Mn2+, Cu2+, Mg2+, Hg2+, Al3+, Zn2+, Cd2+, Ni2+, Co2+, Sn2+) and it demonstrated a selective “turn-off” fluorescence quenching for Cu2+ ions, attributing to coordination with the imine (C=N-) nitrogen as supported by DFT studies showing a reduced HOMO-LUMO gap (0.03 vs. 0.159 eV for free probe). Job’s plot confirmed a 1:1 binding stoichiometry between P12 and Cu2+. The sensor exhibited an LOD of 2.4 × 10−8 M and a high Ka value (5.1 × 1011 M−1), revealing its sensitivity and strong binding affinity. P12 also displayed an AIEE, with a 7.47-fold fluorescence increase in a 70% water-THF system. However, Cu2+ sensing by P12 was effective only in non-aggregated states, as aggregation hindered metal interaction. The authors tested the probe for practical applications using real water samples with 94–102% recovery and reversible sensing using EDTA. TLC-based visualization under UV light enabled on-site detection.

Figure 17.

Synthesis of bis-carbazole-based probe P12.

Copper detection has been widely explored using carbazole-based Schiff bases. Another article by Hu and Tao et al. [36] presented the design and synthesis of an AIE-active fluorescent probe (P13) utilizing benzothiazole-Schiff base framework integrated with a carbazole donor group for selective Cu2+ detection in aqueous media. P13 was synthesized via a multistep route involving a Schiff base reaction between carbazole-benzothiazole intermediate (17) and a salicylaldehyde-hydrazone (18), yielding a D-π-A structure in the P13 (Figure 18). Characterization by NMR and IR confirmed the structure. The probe demonstrated high selectivity for Cu2+ over 20 competing metal ions, with minimal Fe3+ interference, and functioned effectively across a broad pH range (4–11). The probe operated via an ESIPT mechanism, exhibiting strong fluorescence at 553 nm in aggregated states (THF/H2O, 1:9 v/v).

Figure 18.

Synthetic route for P13, and binding mode with Cu2+.

Upon Cu2+ binding, chelation-enhanced quenching (CHEQ) occurred due to the paramagnetic nature of Cu2+, resulting in a “turn-off” response. Job’s plot analysis revealed a 1:2 binding stoichiometry (P13:Cu2+), supported by IR and 1H-NMR studies showing coordination of one copper ion through deprotonated hydroxyl groups and imine nitrogen of hydrazine, while the other copper bound to oxygen and nitrogen of benzothiazole moiety, as shown in Figure 18. Fluorescence titration established a linear detection range of 0–45 μM, a low detection limit (LOD) of 0.2 μM (3σ/k method), and a binding constant (Ka) of 1.59 × 104 M−1 via the Benesi-Hildebrand equation. Real-sample testing in environmental water (Dongting Lake, tap water) showed recovery up to 99.78–101.28%. Practical applications were demonstrated using probe-loaded filter paper strips and silica gel, enabling rapid visual Cu2+ detection under UV light. The probe’s reversibility with EDTA and rapid response time (10 seconds) further underscore its potential for real-time environmental monitoring, meeting WHO guidelines for copper detection in drinking water.

In recent years, large biological molecules, specifically steroids, have attracted great attention in ion recognition, due to their amphiphilic and structurally rigid conformations. Carbazole and cholic acid naturally exhibit compatibility, having a rigid structure. Additionally, the hydroxyl groups present in cholic acid can facilitate subsequent derivatization, and sensors based on cholic acid could be used under physiological conditions due to biocompatibility. In such a report by Mishra et al. [37] an AIEE-active fluorescent probe, based on cholyl hydrazide carbazole-Schiff base (P14), derived from cholic acid and carbazole for selective detection of Cu2+ ions, was described. The probe P14 was synthesized by refluxing cholylhydrazide (19) (derived from ethyl cholate and hydrazine) with 16 in ethanol, yielding 70% product (Figure 19). Structural confirmation was achieved via FT-IR, 1H/13C-NMR, and photophysical analyses (Figure 20).

Figure 19.

Synthesis of P14.

Figure 20.

Binding mode of P14 with Cu2+.

P14 exhibited AIEE properties in a 70% water-THF mixture, showing a 5.93-fold fluorescence enhancement compared to pure THF, attributed to restricted intramolecular rotations in aggregated states. Dynamic light scattering (DLS) and FE-SEM revealed reduced particle size (139 nm) in aggregated states, correlating with emission enhancement. The probe demonstrated selective “turn-off” fluorescence quenching for Cu2+via PET by coordinating from the imine nitrogen and carbonyl oxygen with minimal interference from other cations/anions tested. Job’s plot confirmed a 1:1 binding stoichiometry, while Stern-Volmer analysis yielded a high binding constant (Ka = 5.3 × 107 M−1) and a low detection limit (1.59 × 10−7 M). DFT studies revealed a narrowed HOMO-LUMO gap (1.42 vs. 4.13 eV free probe) upon Cu2+ coordination, validating the quenching mechanism. The practical applicability was demonstrated in tap and river water samples, achieving recoveries of 92.2–107.7%. Additionally, P14 functioned as an IMP logic gate with Cu2+ and EDTA as inputs.

Advertisement

3. Detection of heavy metal ions (Hg, Pb)

Detecting heavy metal ions is crucial due to their significant impact on human health, environmental safety, and industrial processes. Many heavy metals, such as lead, mercury, and cadmium, are highly toxic even at low concentrations and can accumulate in living organisms, causing serious diseases. Industrial activities often release these ions into water, soil, and air, making their monitoring essential to prevent ecological damage. Strict regulations on heavy metals in water and food drive the need for sensitive, selective, and rapid detection methods for environmental, healthcare, and technological applications.

3.1 Detection of Hg2+

Mercury is a pervasive industrial contaminant used in sectors such as petroleum refining, chemical industry, pharmaceuticals, pulp papermaking, and electronics. It bioaccumulates in brain tissues, leading to neurological damage once critical concentrations are reached. Thiourea moiety is widely used in the field of molecular recognition, especially for heavy metal ions, because of its three coordinating sites, two nitrogen atoms, and one sulfur atom (H2NCSNH2). The sulfur group has a special affinity for mercury, as the name mercapto for -SH group comes from the Latin words “mercurium captans,” meaning capturing mercury. Combining carbazole with thiourea could elevate its ability to capture Hg metal ions. Li et al. [38] proposed a carbazole-Schiff base (P15) synthesized from 20 and 4-phenylthiosemicarbazide (21) in ethanol, as shown in Figure 21. The pale-yellow compound obtained was tested by UV-Vis absorption spectroscopy for a variety of metal ions (Na+, Ag+, Ca2+, Mg2+, Zn2+, Cu2+, Fe2+, Fe3+, K+, Cd2+, Cr3+, Co2+, Mn2+, Pb2+, and Ni2+), in DMF. The compound showed two absorption bands in the UV spectrum at 290 and 359 nm. P15 specifically showed a hyperchromic shift for band at 290 nm and a large red shift or bathochromic shift from 359 to 385 nm for Hg2+ ion, while the color of the solution turned yellow from colorless with no obvious response for other metals, suggesting it is a colorimetric chemosensor for Hg2+. Further fluorescence experiments suggested that the Hg2+/P15 ratio was 1:2 during binding. The fluorescence emission from 480 to 490 nm showed a red shift only for Hg2+ ion, while for other metal ions, no significant change was observed. Further experiments exhibited an association constant Ka = 2.8 × 104, and a detection limit of 6.5 × 10−6 mol/L as determined by Benesi-Hildebrand, Stern-Volmer, and Job’s plot analyses.

Figure 21.

Synthesis and binding mode of P15 toward Hg2+.

Replacing amino groups with hydrazine and including heterocyclic aromatics not only increases the power of the coordination site but also the extension of π-conjugation. The article by Sekar et al. [39] presented the synthesis and application of a carbazole-hydrazinobenzothiazole-based fluorescent sensor (P16) for selective Hg2+ detection. P16 was synthesized via condensation of 9-ethyl-9H-carbazolyl-3-carbaldehyde (22) and 15 in methanol (Figure 22), yielding a 95% product characterized by NMR, IR, and mass spectrometry.

Figure 22.

Synthesis of P16.

The sensor exhibited a turn-on fluorescence response at 433 nm upon Hg2+ binding, driven by an ICT mechanism from the carbazole donor to the benzothiazole acceptor. UV-Vis and fluorescence titration confirmed selectivity for Hg2+ over 16 competing metal ions, with a 1:1 binding stoichiometry (Job’s plot) and a binding constant Ka = 1.88 × 105 M−1via the Benesi-Hildebrand equation. The coordination involved the imine nitrogen (‒C=N‒) and benzothiazole sulfur, validated by NMR peak shifts (NH proton disappearance at δ 12.17 ppm upon complexation) and IR spectral changes (shifting of C=N and C‒S stretches). The sensor demonstrated a low LOD of 1.41 × 10−7 M fluorometrically and 9.2 × 10−8 M electrochemically, in DMSO/H2O (1:9) at pH 6–10. Protein binding studies with bovine serum albumin (BSA) revealed strong interaction (Ka = 3.14 × 106 M−1), inducing conformational changes in tryptophan/tyrosine residues, as shown by 3D fluorescence. Cyclic voltammetry highlighted redox potential shifts for Hg2+ detection, emphasizing dual optical-electrochemical utility. The sensor’s stability and applicability in biological and environmental monitoring showed its potential for Hg2+ sensing in complex matrices (Figure 23).

Figure 23.

Fluorescence spectrum and binding mechanism of probe P16.

The thiol (-SH) group can be replaced by pyridine/quinoline nitrogen in the vicinity of methylenehydrazine to serve as a coordinating site for Hg2+ ions. The advantage of introducing quinoline moiety is to enhance the fluorescence capacity and quantum yield of designed probe. To this approach, Singh et al. [40] have reported a carbazole-quinoline tagged fluorophore P17, which was synthesized by condensing 22 and 2-hydrazinoquinoline (23) in methanol with a catalytic amount of glacial acetic acid under reflux for 1 h, followed by slow evaporation to yield yellow crystals (Figure 24).

Figure 24.

Synthetic route to probe P17 and its binding mechanism with Hg2+.

Single-crystal XRD analysis revealed that P17 crystallized in a tetragonal I41/a space group, having an almost planar molecular conformation stabilized by strong intramolecular hydrogen bonding (2.63 Å) and intermolecular N–H···N networks. In CH3CN/H2O (1:1, v/v), it exhibited π–π* absorption bands at 330 and 370 nm that diminish markedly upon the addition of Hg2+, accompanied by a color change from pale yellow to colorless in comparison with other 15 metals tested, including Pb and Bi. Its bright blue fluorescence emission at 485 nm underwent complete quenching, and a blue shift to 475 nm upon Hg2+ coordination was observed, while other competing metal ions elicit a negligible response.

Fluorescence titration and Job’s plot analyses confirmed a 1:1 P17-Hg2+ stoichiometry with Ka = 1.01 × 105 M−1 value. The LOD was 2.59 × 10−8 M. Mechanistic investigations combining FT-IR, ESI-MS, 1H-NMR titration, and DFT calculations revealed bidentate coordination through azomethine-N and quinoline-N with concomitant –NH deprotonation, leading to a photo-induced electron transfer (PET) pathway (Figure 24). The authors tested probe P17 for practical applicability by using P17-coated paper strips for on-site Hg2+ detection and successful analysis of lake and Ganga river water samples with recoveries of 94–99%.

3.2 Detection of Pb2+

Recently, Schiff base formation has also become a foundation reaction for making COFs. Carbazole remains at the core of COFs materials, where the Schiff base reaction is being used to construct these 2-D or 3-D materials. The excellent crystallinity, porosity, and tunable pore size of COFs are among the ideal properties for metal ions detection and adsorption.

A carbazole-grafted covalent organic framework (COF-CB) was synthesized by Wang et al. [41] initially by condensation of 1,3,5-tris(4-aminophenyl)benzene and 2,5-dihydroxy-1,4-benzenedicarboxaldehyde to form COF-DhaTab, followed by post-grafting carbazole units onto the framework via Williamson ether reaction as shown in Figure 25. The resulting COF-CB exhibited an eclipsed AA-stacking structure with microporosity (3 nm pore size) and retained crystallinity after functionalization. COF-CB demonstrated a selective “turn-on” fluorescence response to Pb2+ ions via ICT upon coordination of Pb2+ with nitrogen (imine groups) and oxygen (phenolic hydroxyl) sites within the framework. This coordination suppressed non-radiative decay, enhancing fluorescence emission at 540 nm. The sensor showed a linear detection range of 0–80 μM Pb2+ with a low detection limit (LOD) of 1.48 μM (calculated via 3σ/s method) and excellent anti-interference against 15 competing metal ions (Zn2+, Cr3+, Fe2+, Fe3+, Hg2+, Al3+, Cu2+, Ni2+, Mg2+, Ca2+, Mn2+, Ba2+, Co2+, Na+, and K)+. Binding stoichiometry was inferred as 1:1 (Pb2+:COF-CB) based on Job’s plot analysis of similar systems, though explicit stoichiometric data were not provided. Reversibility was confirmed using EDTA, which quenched fluorescence by displacing Pb2+. The COF-CB’s selectivity stemmed from its tailored pore structure and coordination sites, enabling Pb2+ detection in real water samples (tap/drinking water) with 95–98% recovery. While the association constant (Ka) was not explicitly reported, the high sensitivity and selectivity underscored strong binding affinity. This work highlights COF-CB as a robust, reusable chemosensor for environmental Pb2+ monitoring, leveraging its structural stability and fluorescence enhancement mechanism.

Figure 25.

Structure of carbazole-based COF and binding mechanism with Pb2+.

The above one is just an example of carbazole-based Schiff base COFs materials’ potential toward metal ion detection; there are many exciting carbazole-based Schiff base COFs, for the selective and sensitive detection of metal ions. Since this topic deviates from the title of chapter so COFs are not discussed here.

Advertisement

4. Conclusions

So, the bottom line is that carbazole-based Schiff-base chemosensors unite synthetic simplicity with versatile photo-physics. By tuning the donor/acceptor balance, harnessing PET/CHEF/AIE/ESIPT mechanisms, and integrating user-friendly formats (gels, strips, smartphone read-outs), detection limits can be pushed into the nM range, to achieve rapid “turn-on” selectivity in real water samples. For common transition metals such as Fe, Cr, Cu, Al, etc., imine nitrogen (C=N) alongside other donor groups like -OH, -SH, NH2, NR2, etc. are ideal as coordinating sites while incorporating carbazole and naphthalene moieties enhance photophysical and fluorescence properties in the designed chemosensor. For heavy metals (Hg, Pb, As), the utilization of carbazole-based COFs and MOFs and nanomaterials especially nanofibers could be designed to detect such metals in the nM range. As global concerns over metal pollution and metabolic disorders escalate, carbazole-Schiff bases stand poised to play a pivotal role in advancing affordable, rapid, and reliable detection technologies.

Advertisement

5. Future perspectives

Although carbazole-Schiff bases could be promising candidates for developing chemosensors for a variety of sensing applications, because of their ease of formation and stapling of two moieties of tunable characteristics together, their labile hydrolysis and water sensitivity remains a challenge. In addition, the Lewis acidic behavior and catalysis ability of transition metal ions promote the Schiff bases hydrolysis in both aqueous and organic/aqueous solvents. However, a strategically designed configuration of binding sites and signaling gears holds the potential to yield a chemosensor capable of detecting the metal ions in aqueous or organic/aqueous media. So next-generation carbazole-based Schiff base sensors can utilize a push-pull architecture by combining strong donors (dialkylamino) with extended acceptors (dicyanovinyl, nitrophenyl) to red-shift absorption/emission toward the visible/NIR. To overcome water solubility problem sulfonate or poly(ethylene glycol) moieties can be introduced to eliminate organic cosolvent for sensing in full aqueous media. Integration of emerging techniques like nanostructured assemblies the sensor materials can be imbedded in paper strips, hydrogels, or nanofibers for simple dip-test, on-site colorimetry, or “dipstick” fluorescence methods.

Despite these advances, translating laboratory-scale discoveries into commercial sensors remains a challenge. Key hurdles include long-term stability under varying pH and temperature conditions, reducing synthesis costs, and integrating these materials into user-friendly devices such as paper-based strips or portable fluorimeters. However, progress in nanotechnology and materials engineering, such as embedding carbazole-Schiff bases into polymer matrices or metal-organic frameworks (MOFs) offers promising pathways to enhance durability and scalability. Additionally, collaborations between academia and industry are critical to validate sensor performance in real-world scenarios, from monitoring industrial effluents to diagnosing metal imbalances in clinical settings.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Jia Y, Li J. Molecular assembly of Schiff base interactions: Construction and application. Chemical Reviews. 2015;115(3):1597-1621
  2. 2. Faridbod F, Ganjali MR, Dinarvand R, Norouzi P, Riahi S. Schiff's bases and crown ethers as supramolecular sensing materials in the construction of potentiometric membrane sensors. Sensors. 2008;8(3):1645-1703
  3. 3. Alorabi AQ, Abdelbaset M, Zabin SA. Colorimetric detection of multiple metal ions using Schiff base 1-(2-thiophenylimino)-4-(N-dimethyl) benzene. Chemosensors. 2019;8(1):1
  4. 4. Reimann MJ, Salmon DR, Horton JT, Gier EC, Jefferies LR. Water-soluble sulfonate Schiff-base ligands as fluorescent detectors for metal ions in drinking water and biological systems. ACS Omega. 2019;4(2):2874-2882
  5. 5. Singh A, Barman P. Recent advances in Schiff base ruthenium metal complexes: Synthesis and applications. Topics in Current Chemistry. 2021;379:1-71
  6. 6. Wang L, Jiao S, Zhang W, Liu Y, Yu G. Synthesis, structure, optoelectronic properties of novel zinc Schiff-base complexes. Chinese Science Bulletin. 2013;58:2733-2740
  7. 7. Sıdır İ, Sıdır YG, Berber H, Demiray F. Electronic structure and optical properties of Schiff base hydrazone derivatives by solution technique for optoelectronic devices: Synthesis, experiment and quantum chemical investigation. Journal of Molecular Structure. 2019;1176:31-46
  8. 8. Amith Nayak PH, Bhojya Naik HS, Teja HB, Kirthan BR, Viswanath R. Synthesis and opto-electronic properties of green light emitting metal Schiff base complexes. Molecular Crystals and Liquid Crystals. 2021;722(1):67-75
  9. 9. Sakthivel A, Jeyasubramanian K, Thangagiri B, Raja JD. Recent advances in Schiff base metal complexes derived from 4-aminoantipyrine derivatives and their potential applications. Journal of Molecular Structure. 2020;1222:128885
  10. 10. Patil SK, Das D. Substituent-controlled selective and sensitive potential optical fluoride sensors based on salicylidene Schiff base derivatives. ChemistrySelect. 2017;2(21):6178-6186
  11. 11. Zhang M, Gong L, Sun C, Li W, Chang Z, Qi D. A new fluorescent-colorimetric chemosensor based on a Schiff base for detecting Cr3+, Cu2+, Fe3+ and Al3+ ions. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2019;214:7-13
  12. 12. Duarte F, Dobrikov G, Kurutos A, Santos HM, Fernández-Lodeiro J, Capelo-Martinez JL, et al. Enhancing water sensing via aggregation-induced emission (AIE) and solvatofluorochromic studies using two new dansyl derivatives containing a disulfide bound: Pollutant metal ions detection and preparation of water-soluble fluorescent polymeric particles. Dyes and Pigments. 2023;218:111428
  13. 13. RasulKhan B, Ponnaiah SK, Balasubramanian J, Periakaruppan P. Novel carbon quantum dotted reduced graphene oxide nanosheets for nano-molar range cadmium quantification. Electrocatalysis. 2022;13(4):435-446
  14. 14. Musikavanhu B, Liang Y, Xue Z, Feng L, Zhao L. Strategies for improving selectivity and sensitivity of Schiff base fluorescent chemosensors for toxic and heavy metals. Molecules. 2023;28(19):6960
  15. 15. Alam MZ, Alimuddin, Khan SA. A review on Schiff base as a versatile fluorescent chemo-sensors tool for detection of Cu2+ and Fe3+ metal ion. Journal of Fluorescence. 2023;33(4):1241-1272
  16. 16. Goshisht MK, Patra GK, Tripathi N. Fluorescent Schiff base sensors as a versatile tool for metal ion detection: Strategies, mechanistic insights, and applications. Materials Advances. 2022;3(6):2612-2669
  17. 17. Taha A, Farooq N, Singh N, Hashmi AA. Recent developments in Schiff base centered optical and chemical sensors for metal ion recognition. Journal of Molecular Liquids. 2024;401:124678
  18. 18. Afrin A, Swamy PCA. Novel Schiff base derivatives for the detection of one-to-multi metal ions and tracking the live cell imaging. Coordination Chemistry Reviews. 2023;494:215327
  19. 19. Zhu W, Yang L, Fang M, Wu Z, Zhang Q, Yin F, et al. New carbazole-based Schiff base: Colorimetric chemosensor for Fe3+ and fluorescent turn-on chemosensor for Fe3+ and Cr3+. Journal of Luminescence. 2015;158:38-43
  20. 20. He Y, Yin J, Wang G. New selective “on-off” fluorescence chemosensor based on carbazole Schiff base for Fe3+ detection. Chemistry of Heterocyclic Compounds. 2018;54:146-152
  21. 21. David CI, Prabakaran G, Karuppasamy A, Veetil JC, Kumar RS, Almansour AI, et al. A single carbazole based chemosensor for multiple targets: Sensing of Fe3+ and arginine by fluorimetry and its applications. Journal of Photochemistry and Photobiology A: Chemistry. 2022;425:113693
  22. 22. Inan-Eroglu E, Ayaz A. Is aluminum exposure a risk factor for neurological disorders? Journal of Research in Medical Sciences. 2018;23(1):51
  23. 23. Bryliński Ł, Kostelecka K, Woliński F, Duda P, Góra J, Granat M, et al. Aluminium in the human brain: Routes of penetration, toxicity, and resulting complications. International Journal of Molecular Sciences. 2023;24(8):7228
  24. 24. Klotz K, Weistenhöfer W, Neff F, Hartwig A, van Thriel C, Drexler H. The health effects of aluminum exposure. Deutsches Ärzteblatt International. 2017;114(39):653-659
  25. 25. Kolcu F, Kaya I. Carbazole-based Schiff base: A sensitive fluorescent ‘turn-on’chemosensor for recognition of Al (III) ions in aqueous-alcohol media. Arabian Journal of Chemistry. 2022;15(7):103935
  26. 26. Erdener D, Kaya İ. Synthesis and characterization of a carbazole-based Schiff base capable of detection of Al3+ in organic/aqueous media. Journal of Fluorescence. 2022;32(6):2097-2106
  27. 27. Bhiri F, Chemli M, Sakly N, Bahrouni Y, Wazzan N, Majdoub M. New chromogenic and fluorogenic symmetric carbazole Schiff base-derivative: Synthesis, sensing performances and smartphone-assisted fluorometric assay for selective Al3+ detection in drinking water. Journal of Molecular Structure. 2025;1322:140548
  28. 28. Lv C, Hu B, Tao Y. A novel AIE-active salicylaldehyde-Schiff base probe with carbazole group for Al3+ detection in aqueous solution. Journal of Fluorescence. 2024:1-10
  29. 29. Fieten H, Gill Y, Martin AJ, Concilli M, Dirksen K, van Steenbeek FG, et al. The Menkes and Wilson disease genes counteract in copper toxicosis in Labrador retrievers: A new canine model for copper-metabolism disorders. Disease Models & Mechanisms. 2016;9(1):25-38
  30. 30. Yang L, Zhu W, Fang M, Zhang Q, Li C. A new carbazole-based Schiff-base as fluorescent chemosensor for selective detection of Fe3+ and Cu2+. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2013;109:186-192
  31. 31. Zhu WJ, Wu ZY, Song JM, Li C. Preparation and properties of two new soluble carbazole-containing functional polyacetylenes. Chinese Chemical Letters. 2011;22(9):1017-1020
  32. 32. Liu X, Xu P, Zhao X, Ge J, Huang C, Zhu W, et al. A novel dual-function chemosensor for visual detection of Cu2+ in aqueous solution based on carbazole and its application. Inorganica Chimica Acta. 2019;495:118975
  33. 33. Ghosh S, Dege N, Kumar A, Misra N, Ghosh K. Naphthalimide-linked carbazole derivative: Synthesis, gelation and naked eye detection of Cu2+ and Fe3+-ions under different conditions. Inorganica Chimica Acta. 2024;572:122280
  34. 34. Liu L, Liu X, Guo C, Fang M, Li C, Zhu W. Carbazole-based dual-functional chemosensor: Colorimetric sensor for Co2+ and fluorescent sensor for Cu2+ and its application. Journal of the Chinese Chemical Society. 2021;68(12):2368-2377
  35. 35. Vyas S, Barot YB, Mishra R. Novel anthracene and carbazole based aggregation induced enhanced emission active Schiff base as a selective sensor for Cu2+ ions. Journal of Fluorescence. 2025;35:2903-2915
  36. 36. Lv C, Hu B, Tao Y. A novel AIE-active benzothiazole-Schiff base probe with carbazole group for Cu2+ detection in aqueous solution. Journal of Molecular Structure. 2025;1328:141259
  37. 37. Bariya D, Vyas S, Mishra R, Mishra S. AIEE active Cholyl Hydrazide Carbazole-Schiff base for selective detection of copper (II) ion. Journal of Photochemistry and Photobiology A: Chemistry. 2024;451:115519
  38. 38. Yin FF, Zhu WJ, Fang M, Wu ZY, Xu Y, Wang HL, et al. A new fluorescence chemosensor for Hg2+ based on carbazole and thiourea. Heterocycles. 2017;94(9):1719-1727
  39. 39. Leslee DBC, Karuppannan S, Kothottil MM. Carbazole-hydrazinobenzothiazole a selective turn-on fluorescent sensor for Hg2+ ions–its protein binding and electrochemical application studies. Journal of Photochemistry and Photobiology A: Chemistry. 2021;415:113303
  40. 40. Singh AK, Singh A, Yadav P, Singh AK, Singh VP. Carbazole-quinoline based ultrasensitive fluorometric sensor for detection of Hg2+ in aqueous medium: Crystal structure, DFT and real sample application. Journal of Molecular Structure. 2025;1337:142197
  41. 41. Zhang N, Wei B, Ma T, Tian Y, Wang G. A carbazole-grafted covalent organic framework as turn-on fluorescence chemosensor for recognition and detection of Pb2+ ions with high selectivity and sensitivity. Journal of Materials Science. 2021;56:11789-11800

Written By

Syeda Aaliya Shehzadi and Mustaghees Ur Rehman

Submitted: 17 May 2025 Reviewed: 21 May 2025 Published: 09 July 2025