Next Article in Journal
Recent Use of Hyaluronic Acid in Dental Medicine
Previous Article in Journal
As-Cast Magnesium Alloys with Ca Addition as a Replacement for Magnesium Alloys with Rare Earth Metals
Previous Article in Special Issue
Effect of Bio-Cementation Level and Rainfall Intensity on Surface Erosion Resistance of Biotreated Slope Using PEICP Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Recent Advances in Carbon-Based Sensors for Food and Medical Packaging Under Transit: A Focus on Humidity, Temperature, Mechanical, and Multifunctional Sensing Technologies—A Systematic Review

1
Centre for Engineering Innovation and Research, School of Engineering, Computing and Mathematical Sciences, Faculty of Science and Engineering, University of Wolverhampton, Wolverhampton WV1 1LY, UK
2
Research Institute of Healthcare Sciences, School of Pharmacy & Life Sciences, Faculty of Science and Engineering, University of Wolverhampton, Wulfruna Street, Wolverhampton WV1 1LY, UK
3
Open Laboratory on Experimental Micro and Nano Mechanics (OLEM), Institute of Mechanics, Bulgarian Academy of Sciences, Acad. G. Bonchev Str. Block 4, 1113 Sofia, Bulgaria
4
Research and Development of Nanomaterials and Nanotechnologies—NanoTech Lab Ltd., Acad. G. Bonchev Str. Block 4, 1113 Sofia, Bulgaria
5
DZP Technologies Limited, Cambridge CB4 2HY, UK
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(8), 1862; https://doi.org/10.3390/ma18081862
Submission received: 1 March 2025 / Revised: 21 March 2025 / Accepted: 3 April 2025 / Published: 18 April 2025

Abstract

:
All carbon-based sensors play a critical role in ensuring the sustainability of smart packaging while enabling real-time monitoring of parameters such as humidity, temperature, pressure, and strain during transit. This systematic review covers the literature between 2013 and 16 November 2024 in the Scopus, Web of Science, IEEE Xplore, and Wiley databases, focusing on carbon-based sensor materials, structural design, and fabrication technologies that contribute to maximizing the sensor performance and scalability with particular emphasis on food and pharmaceutical product packaging applications. After being subjected to the inclusion and exclusion criteria, 164 studies were included in this review. The results show that most humidity sensors are made using graphene oxide (GO), though there is some progress toward cellulose and cellulose-based materials. Graphene and carbon nanotubes (CNTs) are predominant in temperature and mechanical sensors. The application of composites with structural design (e.g., porous and 3D structures) significantly improves sensitivity, long-term stability, and multifunctionality, whereas manufacturing methods such as spray coating and 3D printing further drive production scalability. The transition from metal to carbon-based electrodes could also reduce the cost. However, the scalability, long-term stability, and real-world validation remain challenges to be addressed. Future research should further enhance the performance and scalability of carbon-based sensors through low-energy fabrication techniques and the development of sustainable advanced materials to provide solutions for practical applications in dynamic transportation environments.

1. Introduction

The rapid expansion of global trade and online shopping is coupled with an increased challenge in maintaining product quality, safety, and integrity during storage and transportation. Packaging plays an important role in maintaining product characteristics, especially for food and medical or pharmaceutical products, which are susceptible to contamination, degradation, spoilage, and physical damage leading to health risks, economic losses, and environmental impacts [1,2,3,4]. Indeed, GBP 14 billion of food is wasted annually in the UK [5], and the UK Health Security Agency, UKHSA (2022), reported GBP 5.7 million worth of vaccines wasted in 2019 and that 76% of these losses could have been prevented with better control of logistics, transportation, and storage conditions. At a global level, an estimated 1.3 billion tons of food are wasted throughout the food supply chain [6]. This waste corresponds to approximately 3.3 gigatons of CO2 equivalent emissions each year [7], accounting for 8–10% of total global greenhouse gas emissions [2]. If it were considered as a country, food waste would represent the third biggest source of these emissions worldwide [6]. These statistics underscore the urgent need for advanced packaging solutions to ensure product safety, minimize waste, and promote sustainability. This need is already exemplified by the expected smart packaging market growth from USD 23.33 billion in 2023 to USD 40.02 billion by 2032 [8], with food/beverages having the largest market share (34.6% share) when compared to other segments (Figure 1).
Smart packaging encompasses a range of technologies that integrate embedded sensors, identifiers, and various tools to enhance products safety, efficiency, sustainability, traceability, and user experience. This includes intelligent packaging, which monitors and communicates information about the product quality and state [9], and connected packaging through the Internet of Things and cloud systems [10]. This review focuses on sensors, which allow the monitoring and response of smart packaging to the dynamic transportation environments, where products are exposed to various fluctuating conditions such as humidity, temperature, and mechanical stress, thereby compromising their quality, safety, and integrity.
Packaging for food and medical/pharmaceutical products is designed to accommodate different internal pressures based on the packaging type, such as vacuum-seal packaging (0.1–100 kPa) [11], modified atmosphere packaging (5–50 kPa) [12], flexible or semi-rigid packaging (0.1–50 kPa) [13], and bulk packaging (1–500 kPa) [14]. Additionally, packaging is likely to encounter various mechanical stresses, including vibrations, compression and impacts from drops during storage, handling, and transportation (Table 1). These mechanical stresses can be monitored using mainly pressure and strain sensors. Pressure sensors measure the force per unit area of the material, helping to detect internal pressure changes to identify product leakage, packaging sealing damage, or defects, such as in blisters incomplete sealing and delamination. They also monitor external forces for preventing bursts and supporting the monitoring of packaging integrity under transportation stresses such as stacking, handling, and impacts. Conversely, strain sensors detect the deformation of the material [15] and can find applications in detecting cracks, fatigue, or tears that may jeopardize the packaging barrier properties and compromise the product safety, quality, and economical value.
Humidity sensors play vital roles in maintaining optimal moisture levels, issuing real-time alerts to prevent microbial growth and dehydration that could compromise product texture, potency, or freshness. Similarly, temperature sensors monitor thermal conditions, providing immediate feedback to prevent spoilage, protein denaturation, loss of functionality, and degradation. Multifunctional sensors integrate multiple detection capabilities, such as humidity, temperature, and mechanical sensing, into single platforms.
While acknowledging the key role of smart packaging in ensuring products quality and safety during transportation, reducing waste, extending shelf life and improving supply chain management [10], their general adoption is hindered by challenges related to cost, scalability, and environmental sustainability.
Recent advancements in sensor technology, particularly using carbon-based materials, offer a pathway to overcoming these challenges. These materials including graphene (G), graphene oxide (GO), reduced graphene oxide (rGO), carbon nanotubes (CNTs), carbon nanohorns (CNHs), graphene quantum dots (GQDs), and carbon black (CB), which exhibit exceptional properties, such as high electrical conductivity, mechanical strength and flexibility, thermal stability, light weight, and large surface area [25]. These properties enable enhanced sensing performance, miniaturization, and adaptability to diverse stimuli, including humidity, temperature, pressure, strain, and volatile organic compounds (VOCs). In addition, carbon-based materials are generally more cost-effective and relatively environmentally friendly compared to traditional high-performance materials such as silver [26]. However, certain high-purity carbon-based material production processes, particularly chemical vapor deposition (CVD), involve metal catalysts (nickel, iron, cobalt) which may cause heavy metal contamination due to residual catalyst particles [27]. Effective purification methods, including acid treatments, oxidation, thermal annealing, and magnetic separation, as well as employing low-toxicity or metal-free catalysts, have been developed to remove residual metal catalysts from carbon-based materials [27,28]. Technologies proposing catalyst-free synthesis include mechanical exfoliation, liquid-phase exfoliation, electrochemical exfoliation, Hummers’ process, plasma-enhanced catalyst-free CVD, and laser-induced methods [28,29,30,31] or the use of biomass-derived carbons sources [32], thus, offering environmentally friendly options. These technologies nevertheless still face challenges regarding mass production and production efficiency. To fully harness the environmental and economic potential of carbon-based materials, continued research is essential in developing greener synthesis techniques, effective catalyst management, and comprehensive lifecycle assessments. Although limited study has been performed on the interaction of carbon-based material such as graphene with the environment, it has been suggested that they can be biodegradable [33], therefore, offering the possibility of end-of-life disposal without the need for complex separation and recycling processes, especially when used in composites with other biodegradable polymers such as cellulose-based materials and polylactic acids.
While existing literature has reviewed carbon-based sensors, these reviews have typically focused on single-sensor functionalities (e.g., humidity, temperature, or mechanical sensors) and broadly addressed diverse applications ranging from environmental monitoring to healthcare wearables [34,35,36,37,38,39,40]. Existing reviews lack detailed discussions on novel carbon-based composite materials and advanced fabrication techniques developed specifically for packaging in relation to transportation conditions or transit. Equality, to the best of our knowledge, no previous review has simultaneously examined multiple sensing functionalities tailored to smart packaging for food and medical/pharmaceutical products. Furthermore, recent regulatory changes [41,42,43] coupled with technological advancements, alongside the growth of online shopping, have heightened the demand for specialized reviews on these applications to ensure compliance and efficacy. Consequently, there is a clear gap in the literature and this review significantly contributes to filling that by providing the first comprehensive and systematic analysis of recent advancements (2013–2024) in carbon-based humidity, temperature, mechanical, and multifunctional sensors for smart packaging applications for food and medical/pharmaceutical products that meet the acceptable conditions for safe storage and transportation (Table 1). Our review uniquely highlights the innovations in composite material design, structural optimization, and fabrication techniques that could enhance sensor performance specifically in packaging scenarios, offering new insights that bridge laboratory innovation with practical industrial scalability and regulatory compliance. Finally, the review identifies current research gaps and outlines potential future directions for cost-effective, scalable, and environmentally friendly on-package carbon-based sensors development in the food and pharmaceutical sectors.

2. Materials and Methods

The methodology for this review follows the guidelines of the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) and the Prisma Checklist can be found in Supplementary Materials, Table S1. This offers a comprehensive review to ensure a comprehensive and focused analysis of recent advancements in carbon-based sensors [44]. It was registered in the Open Science Framework (https://doi.org/10.17605/OSF.IO/DYX4H, accessed on 2 April 2025).

2.1. Database Selection and Search Strategy

A comprehensive search was conducted across major academic databases, including Scopus, Web of Science, IEEE Xplore, and Wiley to ensure a broad coverage of relevant studies. The search strings were used to capture articles in all databases using a structured combination of keywords such as TITLE-ABS-KEY (‘carbon-based’ OR ‘graphene-based’ OR ‘graphene’ OR ‘carbon nanotube’ OR ‘graphene oxide’ OR ‘CNT’ OR ‘carbon’) AND TITLE-ABS-KEY (‘temperature sensor’ OR ‘humidity sensor’ OR ‘mechanical sensor’) AND TITLE-ABS-KEY (packaging AND food OR medical) AND PUBYEAR > 2013 AND PUBYEAR < 2025 AND (LIMIT-TO (DOCTYPE, ‘ar’)) AND (LIMIT-TO (LANGUAGE, ‘English’)).

2.2. Screening

After initial searching, a multi-step screening process was conducted. Firstly, after removing duplicates, each study was screened by titles and abstracts to access their potential relevance based on predefined inclusion and exclusion criteria (Table 2).
Articles that met the inclusion criteria underwent a detailed full-text review to ensure alignment with the research objectives. To ensure comprehensiveness, citation screening was employed during the full-text review. References within selected articles were assessed to identify additional relevant articles not captured in the database search. Newly identified articles were subjected to the same screening and inclusion criteria.

2.3. Data Extraction

Data extraction was conducted by S.G. using Microsoft Excel for data synthesis and presentation. For each study, the critical information like carbon-based material type, sensor design, fabrication methods, and performance metrics (e.g., sensitivity, response time, operating range, and durability) were recorded and compiled into tables. Owing to the heterogeneity of study designs, only a qualitative assessment was employed.

3. Results

A total of 953 articles were screened, and after excluding duplicates and studies that did not meet the eligibility criteria, 163 relevant articles were identified. The complete selection process is shown in Figure 2. The PRISMA flow diagram illustrates the systematic review process, detailing the number of articles included and excluded at each stage. Among the relevant articles, 72 deal with humidity sensors, 45 with temperature sensors, 23 with mechanical sensors (10 pressure sensors and 13 strain sensors), and 23 articles related to carbon-based multifunctional sensors. Articles included in this review are largely from Journals of Quartiles, Q1 and Q2.

3.1. Humidity Sensors

Table 3 summarizes 72 carbon-based humidity sensors with potential application in packaging for food and medical/pharmaceutical products, detailing the materials used, production methods, and resulting properties.
Among the carbon materials studied, GO-based sensors were the most commonly used (33 studies) [45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77], followed by graphene (17 studies) [78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94], rGO (8 studies) [95,96,97,98,99,100,101,102,103], CNTs (7 studies) [68,76,104,105,106,107,108] and GQDs (4 studies) [109,110,111,112]. Sensor performance is frequently enhanced by incorporating carbon-based materials with polymers, metal oxides, specific dopants, and advanced nanostructural materials. A shift is noted from metal- to carbon-based electrodes, such as graphene [67,84,85], GO [77], rGO [97], Laser induced graphene (LIG) [53,54,62,78], carbon [113], CNTs [72], and their composites [77], to support cost reduction and improve flexibility and durability.
The fabrication techniques are diverse and tailored to different sensor configurations and specific performance. Key methods include drop-casting, spin-coating, and screen-printing created uniform films, whereas electrospinning, laser scribing, hydrothermal techniques form 3D or textured structures. The main types of investigated humidity sensors are resistive (29 studies), capacitive (18 studies), and impedance (11 studies), with a growing interest in hybrid configurations, which combine different sensing mechanisms for improved environmental adaptability.
Many carbon-based humidity sensors operate effectively across 10–97%RH, and some sensors like oxidized carbon nanohorns/graphene oxides/Tin Oxide/Poly(vinylpyrrolidone) (CNH/GO/SnO2/PVP) [47], GO/oxidized CNH/PVP [48], and molybdenum ditelluride/graphene (MoS2/graphene) [78], extended to 0–100%RH detection. In contrast, several sensors, such as N-S co-doped GQDs, have a narrow range (40–90%RH) with optimized performance [110]. Sensitivity varies significantly based on material composition and structure, ranging from 0.022 for ZnO/PVP-rGO [96] to 9,750,000% for paper cellulose fiber/GO [77].
The response and recovery time is essential for real-time monitoring and the values vary from 0.02 s for nanocrystalline graphite [114] to 333 s for TEMPO-oxidized cellulose fibers/carbon nanotubes (TOCFs/CNTs) [106]. Capacitive sensors generally show faster response times than resistive and impedance sensors. Notably, 18 sensors, including ZnO/PVP-rGO nanocomposite demonstrated faster recovery than response time [53,54,55,57,58,59,63,64,70,71,72,75,77,78,82,96,101,111].
The selectivity of carbon-based sensors is key in ensuring reliable sensor performance under real-world conditions. The reviewed studies [46,52,55,59,62,88,113] demonstrated high selectivity to water vapor through material modifications such as doping, composite integration, and surface functionalization. The reusability, inferred from stable performance under repeated humidity cycles, is indirectly supported by the stability and durability data, though direct cyclic reuse metrics have not been reported.
Sensor stability is crucial for consistent sensor performance in different transportation scenarios of products whereas low hysteresis is important for its long-term reliability. The stability was up to 1095 days for laser-reduced GO/MWCNT sensors, fabricated using a 785 nm, 5 mW laser with a 50 µm spot size under Direct Laser-Scribed (DLS) conditions [97]. The reliability remains around 3–8% for most sensors, with the exception of Li-doped GO achieving 0.83% [45]. Some sensors, such those with GO functionalized with hydroxyl groups [46], shellac-derived carbon thin film [113], oxidized CNH/GO/SnO2/PVP nanocomposite [47], and GO-oxidized CNH-PVP [48] consuming 15 µW [46] to 2 mW, could be ideal for prolonged monitoring. This is a very useful feature as low power consumption sensors are advantageous in many respects from a sustainability and environmental standpoint as well as for system miniaturization.
Table 3. Summary of different humidity sensors reported in the literature with their resulting properties including the sensing range, sensitivity, durability, linearity, response, and recovery time.
Table 3. Summary of different humidity sensors reported in the literature with their resulting properties including the sensing range, sensitivity, durability, linearity, response, and recovery time.
MaterialFabrication TechniqueTypeSensing Range (%RH)SensitivityResponse/
Recovery Time (s)
Stability (Days)LinearityRemarks/CommentsRef
Li-doped GODrop castingResistive11–9717.13–3038.16% 4/25 Not reported (N)YesHysteresis is 0.83% and thermal stability is 850 °C.[45]
GO functionalized with hydroxyl groups6–95~38.58.5/13390 Yes Hysteresis is 0.63%.
High selectivity to humidity.
Power consumption is 15 µW.
[46]
Oxidized CNH/GO/SnO2/PVP nanocomposite film0–1000.9021 Ω/% RH42/164NYes CNHox/GO/SnO2/PVP mass ratio is 1/1/1/1.
Power consumption is <2 mW.
[47]
GO-oxidized CNH-PVP0–1000.15–0.240–90/62–73NYesOptimal GO:CNH:PVP is 1:1:1.
Power consumption is <2 mW.
[48]
Ultra-thin, single-layer GO film10%–95120.57%/%RH0.49/0.6560 NoOptimal sensor has 300 nm GO with 20 μm electrodes spacing.[49]
Oxidized single-walled carbon nanohorns (SWCNHs)10–90~2.1 × 107 Ω/RH (air)3/N (air)NYesSurface area is 1300–1400 m2/g.[115]
~9.1 × 106 Ω/RH (N2)8/N (N2)
GO/PVA compositeResistive
Frequency
20–80−12,000 Ω/%RH NNNIt achieves ~1.8% RH resolution.[50]
0.0001 kHz/%RH
rGO/PVDF compositeSolution castingResistive11–9798.99%21/2690 Yes Optimal is 30 vol% rGO/PVDF.
Hysteresis is 5.5% and decomposition from 434° C.
[95]
Endohedral lithium-doped SWCNT/sodium dodecylbenzenesulfonate (Li@SWCNT/SDBS)Arc discharge and drop casting11–974%/%RHNNNoOptimal sensor is five-layer thin film (~5 µm thickness).
Hysteresis is 4.3%.
[104]
GQDs/Ag nanoparticles (AgNPs)Hydrothermal and drop casting25–9598.14%15/15NNoOptimal GQDs/AgNPs is 1:1.[109]
GO filmDrop castingCapacitive 15–9537,800%10.5/4130 YesHysteresis is ~5%.[51]
GO/Ag composite11–9725,809 pF/%RH~8/~1230 NOptimal Ag content is 2 wt%.
Good selectivity for H2O vapor.
[52]
GO0–971800 pF/% RH16/9NYes Spiral LIG as electrodes.
Optimal GO thickness is 50 nm.
3.03% hysteresis.
[53]
GO10–903862 pF/%RH58/1542 NHysteresis is 1.2%.
Optimal sensor used 60 µL GO and 150 µm gap size for LIG interdigitated electrodes (IDE).
[54]
GO/MoTe2 composite nanosheets11.3–97.394.12 pF/%RH39/1235 NOptimal GO to MoTe2 ratio is 1:2.
High humidity selectivity.
[55]
ZnO/PVP-rGO nanocomposite15–95~0.022~12/~387 Yes [96]
GO-Mn-doped ZnO nanocompositeCapacitive 10–90N4.5/2130 Yes95.7 times higher sensitivity in capacitance and 97 times in resistance compared to conventional GO.[56]
Resistive
GO-doped P(VDF-TrFE)/LiCl compositeCapacitive change25–951708.8 pF/%RH7.8/4.5 NYes Pores from 300 nm to 1.1 µm.
Reduced hysteresis due to GO and LiCl modification.
[57]
GOResonant frequency10–900.719 kHz/%RH<78/54 30 NoResolution (0.4% RH), hysteresis (<4%), and minimal response to CO2.[58]
HGO/GO/Mg2+ composite membrane11–970.0343 kHz/%RH7/610 YesHysteresis is ~3.2% RH.
High humidity selectivity.
[59]
GOVoltage33–981.1–10.0 mV/%RH0.28/0.32.5 YesGO thickness is 10 μm.[61]
2D MoS2/graphene nanocomposite foamImpedance0–10050,000–
385,000 Ω/%RH
4/2NNoSensor used LIG as electrodes.
Hysteresis is 3.8%.
[78]
Laser-reduced GO/MWCNT Drop casting and direct laser scribingImpedance11–97350,000 Ω/%RH0.061/2.3 1095 YesSensor used rGO IDE.
Hysteresis is 3.1%.
[97]
Capacitance798 pF/%RHc
Thermally reduced GOSpin-coatingResistive32–65 5%35/N NYesHighly thermal-reduced GO has the optimal performance.[98]
P(VDF-TrFE) with graphene flower compositeCapacitance
Impedance
8–980.0558 pF/% RH0.8/2.515 YesN[79]
GOImpedance6–97182,068.791/%RH0.8/0.91YesTi3C2Tx MXene-based sensor exhibited faster response than sensors using metallic electrodes.[60]
N-S co-doped GQDsHydrothermal and spin-coating40–90N15/5590 NOptimal GQDs content is 10 mg with 2.2% hysteresis.[110]
GQDs/carbon nitride (g-C3N4) composite7–97100,000 Ω/RH44 /10 NYesLow hysteresis (<1%) and high surface area (545 m2/g)[111]
Bi-layered PVA/graphene flower composite filmSpin-coating and spray-coatingCapacitance40–9029,000 pF/%RH2/3.515 NUniform dispersion of PVA/GF layer with ~2.32 µm thickness.[80]
Impedance
Shellac-derived carbon (SDC) thin filmSpray coating and thermal annealingResistive0–900.54/% RH0.14/1.728 YesCarbon IDE.
High selective to humidity.
Power consumption is ~1 mW.
[113]
rGO-sodium dodecyl sulfate (SDS) composite filmDrop-coatingResistive25–9511.4143 Ω/% RH (RT)9/1010 YesHysteresis is 0.04852%.[99]
GO11–971.113 Ω/Ω–%RH2/35NYesSensor used 300 nm wrinkled GO film on the LIG electrode.
Hysteresis is 3%.
High humidity selectivity.
[62]
GOQuartz crystal microbalance (QCM)11.3–97.30.1605 kHz/%RH30/5NYesThe study used the finite element analysis software COMSOL Multiphysics.[63]
Polydopamine-coated cellulose nanocrystals/GO nanocomposite (PDA@CNC/GO)Drop-coatingResonance frequency11.3–97.30.05466 kHz/% RH37/521 NOptimal composition is 30 wt% PDA@CNC.
Hysteresis is 4.3% RH.
[64]
Graphene flower/ZnO compositeSol–gel and spray-coatingResistive15–867.7 µA/%RH0.4/4 NNHigh surface area to volume ratio and pore composite.[81]
GO on tilted fiber grating (TFG)Dip-coatingResonance wavelength
Intensity
30–800.0185 nm/%RH0.042/0.115NYesGO thickness is 54 nm.[65]
GO/PVA composite filmIntensity 20–99.90.529 RH (%)147/293NYesN[66]
Graphene–carbon inkScreen printingResistive25–91.712.4 Ω/%RH~31/~8120 NOptimal configuration is single-layer sensor.[82]
G/polypyrrole/carbon black (CB) composite23–92.712.2 Ω/%RH 5/721 NDurability is 100 bending cycles
Single-layer is the most effective configuration.
[83]
Graphite/WO3 nanocomposite11–9712.7–60.8%NNYesOptimal sensor using graphite/WO3 ratio is 1:3, with <1% hysteresis.
120° bending angles.
[116]
Multilayer GOResonance frequency
Backscattered phase
11–980.5°/%RHNNN30 µm GO film and printed graphene antenna electrodes.[67]
Cellulose nanofiber (CNF) and graphene nanoplatelet (GNP) compositeMixing and screen printingResistive30–90240%17/22240 NComposite with 200 mg GNP as electrode. [84]
Graphene inkInkjet printingCapacitive 10–700.03 pF/%RH2.46/2.6310 NOptimal sensor is six-layer graphene film with graphene IDEs.[85]
GO/CNT−OH/Nafion nanocompositeResonance frequency30–95547 kHz/%RH110/1152.08YesHysteresis is 3%.[68]
Functionalized MWCNTs and hydroxyethyl cellulose (HEC) compositeGravure printingResistive20–800.0485/%RH20/35 0.4YesThe optimal FMWCNT concentration is 2.5 wt%.[105]
Carboxymethyl cellulose@graphene (CMC/G) composite3D printed groove moldImpedance11–9597%300/N16 YesOptimal graphene content is 0.16 wt%.[86]
Graphene filmLiquid phase exfoliation and LB assemblyResistive8–955%0.028/0.03NYes The thickness is ~3.4 nm (~ 10 layers).
Flexibility is 10° bending.
[87]
GOSelf-assemblyCapacitive 30–900.00565 pF/% RH180/N14 NOptimal sensor is 2 mg/mL GO with 2.85% hysteresis.[69]
Pyranine modified-rGO compositeOne-step supramolecular assemblyImpedance11–95IL/IH = 6000<2/~6NYesHysteresis is 8% RH.
Stable for 100 cycles.
[100]
TEMPO-oxidized cellulose fibers (TOCFs)/CNTsElectrostatic self-assemblyCurrent11–9587%333/523 90 Yes Optimal TOCFs-to-CNTs ratio is 30:1 with a thickness of 48.2 µm and 7.3% hysteresis.[106]
G with 3D flower-like ZnO compositeHydrothermalImpedance12–90446120/16030 NOptimal G content is 70 wt% with 2.32% hysteresis.
High humidity selectivity.
[88]
PVDF (polyvinylidene fluoride) with 0.5 wt% GElectrospinning Capacitive35–900.0463 pF/%RHNNYes PVDF/G with Ag electrode, showed 21.3 times faster than DHT11.[89]
SnO2/rGO nanocomposite11–9537,491%80/4NNOptimal rGO doping content is 2 wt%.
Durability is 1000 bending cycles.
[101]
BP/G hybridElectrosprayResistive15–7043.40%9/3028 Yes [90]
GOElectrospray deposition Resonant frequency11–971.74%/%RH54–68/12–2230 NoLow thermal noise.
Optimal is 250 MHz sensor.
[70]
Holey-reduced graphene oxide (HRGO)H2O2-etching-reaction-assisted hydrothermalImpedance11–97−0.04317 log Z/%RH<3/2928 Yes Surface area is 274.5 m2/g.
Hysteresis is 2.57%.
[102]
GO Dripping and vacuum heatingCapacitive20–901.77–164.98 pF/% RH10/2NNHysteresis is 1%.[71]
GODripping and coating10–9016.7 pF/%RH0.0208/0.019980 NThe optimal sensor used 1 mg/mL GO and CNTs as electrodes. Hysteresis (<0.44%).[72]
ZnO nanowires and GQDs compositeDrippingResonance frequency30–9040.16 kHz/%RH~30/~35NNoOptimal GQDs content is 2 mg/mL.
30° bending angle.
[112]
SWCNTsVacuum filtrationResistive15–98246.90%290/510NYesOptimal sensor is suspended aligned. SWCNT beams, with 36 μm suspension lengths.[107]
rGO/PANI compositeFiltration0–98580%~70/~139NYes Hysteresis is 3%.
Optimal rGO to PANI ratio is 5%.
[103]
Laser-induced graphene (LIG)Laser Direct Writing (LDW)Capacitive 30–90N8/10NNoThe porous, hair-like LIG pattern was designed with 2-CAD.[91]
Light-scribed GOLaser scribeImpedance7–971.67 × 106 Ω/%RHN1Yes Hysteresis is 0.3–7%.[73]
G/ZrO2 nanocompositeSol–gel12–9040115/206 YesHysteresis is <1.95%.
Optimal is 40 wt% G/ZrO2
[92]
3D graphene foamModified Hummers’ methodResistive0–85.9N0.089/0.189NNEnergy structure of 3DGF model analyzed via CASTEP in Materials Studio 8.0.[93]
Nanocrystalline graphitePlasma-enhanced CVDResistive15–850.0334%/%RH0.02/NNYes Hysteresis is 5%.
It is meandered strip structure.
[114]
SWCNTImmersion 20–8054.7% (s-CNT)40/100NYesHysteresis is 11.45% (semiconducting-CNT) and 0.31% (metallic-CNT).[108]
2.9% (m-CNT)
G/p-aminophenol/poly-2-hydroxyethyl acrylate (G/p-AP/PHEA)In situ free-radical polymerization0–9429%NNNN[94]
Etched GO filmEtchingCapacitive10–1000.000106 pF/% RH1.011/NNNThe study using COMSOL Multiphysics.[74]
Nanofibrillated cellulose (NFC)/GO/PDMS aerogel compositeUltrasonic dispersion and freeze-drying11–976576.41 pF/% RH57/2NNoPorosity is 99.6%.[75]
GO/MWCNTs hybrid on tilted Fiber Bragg Grating (TFBG)Physical precipitationOptical fiber
Amplitude
30–900.377 dB/%RH4/NNYes Hysteresis is 0.7%. [76]
Paper cellulose fiber/GO matrix (PCFGOM)NImpedance10–909,750,000% (1 kHz)1.3 /0.81YesThe sensor used 0.15 w/w% PCFGOM as active layer and 20 w/w% PCFGOM as electrode layers.[77]
Capacitance1,442,500% (10 kHz)

3.2. Temperature Sensors

Table 4 summarizes 45 carbon-based temperature sensors with their composition, production methods, and resulting properties. Graphene-based sensors were predominant among the reported studied (19 studies) [117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135], followed by rGO (11 studies) [123,136,137,138,139,140,141,142,143,144,145], CNTs (9 studies) [146,147,148,149,150,151,152,153,154], and GO (4 studies) [155,156,157,158]. Other carbon materials, including GQDs [159], carbon dots (CDs) [160], and amorphous carbon [161] were reported in only one study, respectively, highlighting their limited exploration in this field. The incorporation of polymers and metals/metal oxides into carbon-based materials has been extensively explored to enhance the performance of temperature sensors. The combination with metal/metal oxides were predominantly applied to rGO [138,141,145] and graphene [128]. In contrast, polymers like Poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS), polydimethylsiloxane (PDMS), and gelatine were widely combined with various carbon materials, including graphene (6 studies) [118,119,120,121,122,135], GO (2 studies) [157,158], rGO (3 studies) [140,143,144], and CNTs (5 studies) [146,147,148,150,151].
Chemical vapor deposition (CVD) was the most dominant method (10 studies) [123,127,128,129,130,131,132,151,152,153], particularly for graphene and CNT sensors. Methods of coating (12 studies) [117,118,119,120,121,122,123,124,136,137,146,155] including spray coating [117,123,124,136], drop casting (5 studies) [138,147,148,155,156], and printing (6 studies) [139,140,141,149,150,157] was widely used in rGO, GO, and composite sensors, offering simplicity and scalability.
Graphene-based sensors fabricated via CVD exhibited the broadest detection ranges, spanning from −266.55 °C [129] to 302 °C [130], quick response time of ~0.030 s in multilayer graphene [129], and high sensitivities, such as 2.15 Ω/°C in micro-fabricated single-layer graphene [127]. Similarly, rGO-based sensors demonstrated wide detection range of −196.15–299.85 °C in rGO sensor [137], with exceptional sensitivity values of up to 1999%/°C for GQDs/rGO/alumina composite [142]. In contrast, GO-based sensors [155,156,157,158] and CNT-based sensors [146,147,148,150,151,152,153] generally operate within narrower ranges, typically starting at 20 °C, limiting their suitability for applications in chill- and cold-chain environments. However, functionalization and advanced fabrication methods have shown potential in improving detection capabilities. For instance, carboxyl-SWCNTs achieved a range of 0–80 °C but showed slow response (176.4 s) and recovery time (316.8 s) [154] whereas CNTs produced using gravure printing demonstrated a broader temperature range (−40–100 °C) and fast response (0.3 s) and recovery time (4 s) [149].
Response times varied significantly from 0.030 s for multilayer graphene produced by CVD [129] to 306 s for uncovered drop-casted GO sensors [155] depending on the material and encapsulation approach. Composite systems often outperformed pure carbon materials in detection range, sensitives, and responsiveness, particularly those combined with metal or metal oxides. For example, rGO/Ag exhibited extended temperature sensing ranges (−60–80 °C), faster response times (0.47 s), and good sensitivity (0.555 Ω/°C) [138].
Long-term stability was assessed in 10 studies [121,126,138,140,143,146,147,152,156,158], rGO/Ag nanocomposite with Parylene encapsulation demonstrating exceptional longevity, maintaining performance over 730 days durability [138].
Table 4. Summary of different temperature sensors reported in the literature with their resulting properties including sensing range, sensitivity, stability, response, and recovery time.
Table 4. Summary of different temperature sensors reported in the literature with their resulting properties including sensing range, sensitivity, stability, response, and recovery time.
MaterialFabrication TechniqueSensing Range (◦C)Sensitivity/TCR
(%/°C)
Response/
Recovery Time (s)
Stability (Days)Remarks/CommentsRef
GODrop casting20–70 822 Ω/°C 306/554
(uncovered)
NEncapsulation: PDMS[155]
Spray coating20–60 N0.525/0.35
(uncovered)
5.18/9.68
(covered)
rGOSpray coating30–1000.6345%/°C1.2/NNEncapsulation: high-temperature transparent insulating tape. [136]
Multilayer graphene ink film30–9043.27 μV/K0.15 /15 NOptimal sensors have 108 nm thickness and provide 300 μV output voltage, and signal-to-noise ratio is 35.[117]
rGOSpin coating−196.15–299.85−0.801–−32.04%/°C 52/285 NOptimal rGO concentration is 3wt%, with 0.1 °C resolution.[137]
MWCNT doped in polyethylene glycol and PU (MWCNT-PEG-PU) nanocomposites25–50~80%N7Optimal MWCNT concentration is 8 wt% and stable 30 bending cycles.[146]
Graphene-coated microfiber (GCM)Coating22–402.1 dB/°CNNMinimum resolution is 0.0005 °C.[118]
Polyaniline/graphene (GPANI) embedded in Polyvinyl Butyral (PVB) composite filmCoating using Mayer rod25–80 °C−1.2%/°CNNSensor also responds to external pressures (0–30 kPa).
Encapsulation: Bezel tape
[119]
Graphene and gelatin nanocompositeBlade coating−13–37−5.3–−23 mV/°C10.4/NNStable for 20 cycles.[120]
Graphene/gelatin nanocomposite−13–37−19 mV/K41.8/N (pristine sensor)2Energy consumption is 8.1 μWh for pristine sensor. [121]
28.9/N (aged sensor)Energy consumption is 8.5 μWh for aged devices.
PU/G NanocompositeIn situ polymerization and dip coating25–606 pm/°CNNThermal stability to 217 °C from 204 °C. [122]
rGOAir brush spray coating0–10045.1%121/NNN[123]
Graphene nanoplatelets (GNP)52%89/N
Plasma-grown graphene (Gpl)Plasma discharge20.5%125/N
Graphene via CVD (Gcvd)CVD27%68/N
GOPost-COMS MEMS
Drop casting
−70–40155.73–58,555.26 pF/°CNot reported (N)30Capacitance sensor.[156]
rGO/Ag nanocompositeUltrasonication and drop casting−60–800.555 Ω/°C0.47/N (cold)730Encapsulation: Parylene.[138]
3.45/N (hot)
CNT/PEDOT:PSS composite Drop casting25–45 −1.97%/°C (initial)N6Encapsulation: PDMS
Optimal CNT/PEDOT ratios is 1:5.
[147]
−2.80%/°C (6 days aging)
CNT and methylcellulose (CNT/MC) compositeSolution casting20–700.2%/°C6.1/3.1 (hot)
5.2/7.2 (ice)
NStable over 480 cycles.[148]
Graphene Nanoribbons (GNRs)Mask spraying or direct handwriting30–80172%
TCR = 1.27%/°C
0.5/0.5 NUsing MWCNT ink electrodes and Scotch tape encapsulation.
0.2 °C resolution and stable 5000 bending cycles.
[124]
CNTGravure printing−40–100−0.4%/°C 0.3/4 NHigh accuracy (±0.5 °C).
Encapsulation: organic and silver.
[149]
GO/PEDOT: PSS compositeMask printing25–100−1.09%/°C18/32 NEncapsulation: Kapton tape.
Stable 1000 bending cycles.
[157]
Functionalized and reduced graphene oxide via sulfonated aromatic diamine (f-rGO)Inkjet printing30–82−0.0164/°C176.4/316.8 N [139]
CNT/PEDOT-PSS composite25–500.31%/°C~39/~196NEncapsulation: translucent polyurethane welding tape.
Stable 1000 cycles bending.
[150]
rGO with alkali lignin 25–1350.59%/°CN180Sensor used meander-shaped rGO as electrode.[140]
rGO/AgAerosol jet printing0–2000.001162–0.001519/°CNNOptimal four layers rGO/Ag.
Stable 1000 bending cycles.
[141]
Porous LIG CO2 laser-induced
Direct laser writing
1–8N16/58N Encapsulation: PDMS.
Stable 200 bending cycles.
[125]
LIGLaser direct writing24–80−0.58%/°CN14Sensors optimized by finite element analysis photothermal model.[126]
Amorphous carbon filmsDC Magnetron Sputtering20–1501.62 mV/°C
TCR = 0.00128/°C
NNDC magnetron sputtered sensors are more stable and practical than ion-beam-deposited sensor.[161]
Micro-fabricated single-layer grapheneCVD10–301.25 Ω/°C
(SiO2/Si substrate)
NNSensor used graphene electrodes and PDMS gasket encapsulation. [127]
2.15 Ω/°C (SiN substrate)
1.90 Ω/°C (suspended graphene substrate)
CNT forest-PDMS composite30–900.55 Ω/°C NNEncapsulation: PDMS.[151]
Graphene and Lithium Niobate (LiNbO3)10–70−0.23 nm/°CNNEncapsulation: PDMS. [128]
Multilayer graphene−266.55–26.85−1 (THS < −243.15 °C) ~0.030/NNSensor made by seven layers of single-layer graphene.[129]
<0.1 (THS > −173.15 °C)
Vertically aligned CNT filmTCVD20–1104.74 μA/°C (air)N30Triple-electrode structure enables long-term sensor operation.[152]
22.72 μA/°C (N2)
MWCNTCVD and wet transfer22–200 0.0033 V/°C
TCR = 0.00103/°C
NN2.7 μm MWCNT sensor had carrier mobility (−28.5574 cm2/Vs). [153]
Single-layer graphene27–3020.00207/°C (27–177 °C)NNResistance is almost unaffected by humidity. [130]
0.00239/°C (177–302 °C)
Suspended few-layer and multilayer graphene25–120 1.07–3.5%/°C NNN[131]
GrapheneCVD and AI sacrificial layer process 25–2002.134 Ω/◦CNNEnhanced 41.93% consistency.
Encapsulation: SiO2 layer.
[132]
GQDs embedded in a rGO/alumina composite filmSol–gel−196.15–26.85
26.85–99.85
−1999%/°C
−0.98%/°C
~0.3/0.8
3.96/6.01
NShort-term stability is 50 cycles.[142]
CNC-assisted carbon dots (CDs)-grafted SrAl2O4: Eu2+, Dy3+ (SAO) phosphors composite filmSol–gel and vacuum filtration−30–1100.257NNShort-term stability is 3.5 cycles.[160]
High-strength metallurgical graphene (HSMG) Modified PMMA-based transfer−253.15–21.85−0.007/°CNNEncapsulation: transparent polymer.[133]
Polyethyleneimine/reduced graphene oxide (PEI/rGO)Spray dipping25–45
0–60
1.3%/°C0.33–0.443/N120Encapsulation: PDMS.
0.1 °C resolution and 500 bending cycle stability.
[143]
GO/PEDOT: PSS micro/nanowiresSoft lithography30–80−0.007599/°C3.5 /13.4 30Optimal GO doping ratio is 13:1.[158]
LIGCO2 laser irradiation30–60−0.04145%/°C30/NNHigh measurement accuracy (±0.15 °C).[134]
PDA-rGO/sodium alginate/polyacrylamide composite organohydrogelSolvent displacement and cross-linking−20–6097.6%/°C (−20–−5 °C) 0.2/0.3NEncapsulation: VHB tape.
Stable over 3 h.
[144]
10.57%/°C (−5–15 °C)
1.45%/°C (15–60 °C)
Star-like rGO/SnO2/Co3O4 compositeFacile wet chemical precipitation25–1250.561%/°CNN [145]
GNP/PDMS nanocompositeThree-roll milling and molding30–800.052–11.7/°CNNThe optimal GNP concentration is 6 wt%. [135]
Carboxyl-SWCNTs N0–80−225 Ω/°CNNEncapsulation: thermos-reversible polymer.
Self-healing 30 bending cycles.
[154]
GQDs/hollow-core fiberN 10–80−0.01375/°CNNN[159]

3.3. Mechanical Sensors

The review revealed 10 carbon-based pressure sensors that could be used to monitor internal and external forces exerted on packaging as well as 13 carbon-based strain sensors. These findings are summarized in Table 5 and Table 6, respectively.
Similarly to temperature sensors, graphene-based sensors dominate the literature (4 studies) [162,163,164,165] while other carbon materials, including GO [166], rGO [167], CNTs [168], carbon-ink [169], and CB [170] were reported in only one study each. A total of 70% of sensors were porous composite materials that mainly harness the high conductivity and mechanical strength of carbon materials alongside the flexibility, stretchability, and durability of polymers. However, some pressure sensors were produced with only graphene films via CVD [162,163]. The sensor production methods range from simple techniques such as molding [164,167,171], dip coating [169], self-assembly [166], and solvent extraction [165] to relatively more advanced fabrication techniques such as laser thermoforming [170], electrospinning and mechanical drawing [168].
The sensing range depends on materials and structure design, spanning from 20 kPa [167] to 20,000 kPa in a graphene N/MEMS mechanical sensor with crossbeam structure [163]. Porous structures such as porous PDMS [171], graphene/PDMS sponge [164], and polyurethane/graphene (PU/G) foams [165] reach up to 500 kPa, while softer materials such as tannic acid-reduced graphene oxide combined with polyvinyl alcohol (TA-rGO/PVA) hydrogel [167] and nitrogen-doped graphene oxide/dopamine/polyaniline (GO/DA/PANI) aerogel [166] are limited to 20–25.48 kPa, which restricts their use to low-pressure environments and modified atmosphere packaging. Film structures like monolayer graphene have narrow sensing range, up to 80 kPa [162].
The sensor sensitivity also varies, with the formulation and production conditions ranging from 0.0259 kPa−1 [169] to 2200 kHz/kPa [164]. High sensitivity was observed at low pressures and this decreased as the pressure increased, particularly in composites with wide detecting range. Indeed, the sensitivity of a graphene/PDMS composite decreases from 2200 kHz/kPa at 0–10 kPa to 37.5 kHz/kPa at 200–500 kPa [164]. The incorporation of carbon black and use of glucose monohydrate to form porous PDMS/graphene composite contributed to improved flexibility and sensitivity (109.4 kPa−1) of the sensor [170]. Fluoropolymers such as Poly (vinylidene fluoride) (PVDF) and its copolymer poly(vinylidene fluoride-trifluoroethylene) (P(VDF-TrFE)) with SWCNTs or MWCNTs have also been used with reasonable sensitivity. However, recycling or reusing PVDF components when they reach the end of their useful lives is particularly challenging, whereas their disposal by incineration poses an environmental issue because of the potential formation of hydrogen fluoride at elevated temperatures [172].
The durability varies from 100 cycles for hydrogels [167] to 10,000 cycles for porous composites like PDMS/graphene [171]. Response and recovery times varied across materials and sensor structures, and graphene/PDMS sponge achieved fastest response time of 7 ms with 60 ms recovery time [164], while porous PDMS with MWCNT/PEDOT electrode showed 1 s of response and recovery time [171].
Table 5. Summary of different pressure sensors reported in the literature with their fabrication methods and resulting properties including the sensing range, sensitivity, durability, response and recovery time.
Table 5. Summary of different pressure sensors reported in the literature with their fabrication methods and resulting properties including the sensing range, sensitivity, durability, response and recovery time.
MaterialFabrication TechniqueSensing Range (kPa)Sensitivity/Gauge Factor (GF)Response/
Recovery Time (s)
Durability (Cycles)Remarks/CommentsRef
Suspended monolayer graphene (G)CVD0–80GF = 6.73 (circular membrane)Not reported (N)Not reported (N)An improved theoretical model was developed to predict GF and confirm their independence of doping concentration and graphene crystallographic orientation.[162]
GF = 3.91 (rectangular membrane)
GraphenePlasma-enhanced
CVD
0–20,0000.03313 mV/V/kPaN35 daysEncapsulation: Si3N4 film.
Error of hysteresis (2.0119%), nonlinear (3.3622%), and repeatability (4.0271%).
[163]
GF = ~1.35
Graphene/PDMS spongeMixing and molding0.005–50037.5–2200 kHz/kPa~0.007/0.06 5000LC technology used for long-distance wireless transmission.
Optimal graphene concentration is 20%.
[164]
Porous PDMSSugar-cube mold0–1200360–1120 kPa−11/<110,000Sensor used MWCNT/PEDOT composite electrode and low-pass filter.[171]
Tannic acid (TA)-rGO/PVA hydrogelSonication, molding via freeze–thaw0–202.2695 kPa−10.67/0.84100Optimal concentration is 2 mg/mL and tensile strength is 440.213 kPa.[167]
Carbon ink-coated filter paperDip coating0.1–1000.0259–0.627 kPa−1N4000N[169]
PU/G foamsSolvent extraction0–5000.05–7.62 kPa−10.81/0.811000Optimal graphene content is 30 wt%.[165]
Nitrogen-doped GO, dopamine, and polyaniline composite aerogel Self-assembly, freeze-drying, and thermal annealing0–25.480.10 kPa−1N150The optimal mass ratio of GO:DA:PANI is 5:2:2, with 1.46% nitrogen.[166]
P(VDF-TrFE) matrix with MWCNTsElectrospinning and mechanical drawing5–50~540 mV/NNNSelf-powered sensor achieved piezoelectric coefficient of 50 pm/V with 98% linearity.[168]
PDMS/CB/graphene nanosheetsLaser thermoforming0–100109.4 kPa−10.079/0.055 5000CB as an endothermic agent and glucose as a porogen.[170]
Table 6 summarizes 13 carbon-based strain sensors with potential application in food and medical packaging-based monitoring during transportation.
Among the 13 carbon-based strain sensors with potential for application in food and medical packaging-based monitoring during transportation/distribution, CNTs (seven studies) [173,174,175,176,177,178,179], rGO (four studies) [167,177,179,180], and graphene (three studies) [181,182,183] are widely used. These materials were often combined with different highly stretchable and durable polymers or elastomers (e.g., PDMS, PVA, PEI, chitosan, and agar) to improve flexibility and robustness. Several sensors were made of functionalized materials, such as carboxyl-functionalized CNTs [176], polyetherimide-rGO [180], and tannic acid-modified rGO [167], which provided unique features such as self-healing and biocompatibility. Carbon-based strain sensors were generally in the form of layered and 3D composite structures employing hybrid and functionalized materials to enhance sensor performance. Multilayer designs, such as few-layer graphene films [181] has shown to improve sensor sensitivity and mechanical stability. Three-dimensional composites including fragmentized rGO sponge [177] and rGO/MWCNT composites [179] broaden the detection range while also enhancing sensors sensitivity.
Simple and relative cheap methods such as single-step Marangoni self-assembly [182], layer-by-layer assembly [180], sonication [183], and solution casting [175] were used to produce strain sensors which demonstrated low-strain detection (up to 10%) with high sensitivity. Advanced techniques, including microelectromechanical system-assisted electrophoretic deposition (EPD) [173], embedded 3D printing [184], direct writing [179], and screen printing [181] enabled the production of sensors with precise control and broader detection range.
Just for other sensor types, the strain-sensing range varies significantly with material and sensor architecture. The values spanned from 2% in the ultrathin graphene film sensor [182] to 1000% in the carboxyl-functionalized CNTs sensor [176]. Polymer-free CNT sensors fabricated via CVD demonstrated a sensing range of 0–42,100 kPa with high gauge factor (1461), suitable for high-strain applications [174]. Composite sensors’ detection range depends on polymers/elastomers flexibility or stretchability as well as the filler concentration. This ranged from 2% for chitosan/graphene [183] to 280% for TA-rGO/PVA [167]. Functionalized materials, such as carboxyl-functionalized CNTs provided high stretchability and self-healing capability, achieving up to 1000% [176]. Layered and network structure CNT/PDMS sensors achieved up to 100% strain [178].
The durability of carbon-based strain sensors varies significantly, from 100 [167] to 10,000 cycles [175]. The CNT/Agar composite sensor was particularly durable, withstanding up to 10,000 cycles, and demonstrated an enhanced strain range and sensitivity due to increased filler concentrations [175]. The TA-rGO/PVA hydrogel strain sensor exhibited quick response and recovery times (670 ms and 840 ms, respectively), but its soft nature limits its durability to 100 cycles [167].
Only four articles [167,175,177,178] reported any response and recovery time and they vary from 20 ms for FGS/AgNPs/SBS composite [177] to 670 ms for TA-rGO/PVA hydrogel sensor [167].
Table 6. Summary of different strain sensors reported in the literature with their fabrication methods and resulting properties including the sensing range, sensitivity, durability, response, and recovery time.
Table 6. Summary of different strain sensors reported in the literature with their fabrication methods and resulting properties including the sensing range, sensitivity, durability, response, and recovery time.
MaterialFabrication TechniqueSensing Range (%)SensitivityResponse/
Recovery Time (ms)
Durability
(Cycles)
Remarks/CommentsRef
Patterned MWCNT/PDMSMicroelectromechanical system-assisted EPD0–1413–120NNSensitivity tailored by MWCNT film thickness and entanglement.
Sensor adapted to an arbitrarily curve surface.
[173]
Polymer-free CNTsHot-wall atmospheric CVD0–42.1 MPa1461NNHigher sensitivity in IDE devices than single-gap electrodes.[174]
CNT/Agar compositeSolution casting0–1180.28160/250 10,000Increasing filler concentration improved strain from 0.8 to 1.1, and stress from 35.2 to 45.8 kPa.[175]
PDMS-TDI (2,4′-Tolylene diisocyanate)-carboxyl-functionalized MWCNTs nanocompositeOne-pot synthesis, ultrasonication, and casting0–10000.65–2.43N1000Sensor had 98.1% self-healing efficiency at 60 °C over 9 h.[176]
Fragmentized rGO sponge (FGS)/AgNPs/polystyrene-butadiene-styrene (SBS) compositeMultiple-step process0–12020.5–1.25 × 10720/N2000Microcrack contributed to sensitivity.
Sensor had 1521 S/cm conductivity and 680% break elongation.
[177]
CNTs/PDMS0.007–1008765/N1500Optimal sensor had network cracks and 15 layers of CNT.[178]
rGO/MWCNTs compositeDirect writing printing10–4018.55 N900N[179]
Carbon greaseEmbedded 3D printing4003.8N1000Up to 10% variation from its original value after large strains.[184]
Few-layer grapheneMechanical exfoliation and screen printing0–620.02 N100,000
flexing cycles
Number of prepared graphene layers was 2–5 layers.[181]
1000
abrasion cycles
TA-rGO/PVA hydrogelSonication and molding via freeze–thaw cycles0–2801.936 78670/840100Optimal concentration is 2 mg/mL and tensile strength is 440.213 kPa.[167]
PEI-rGO nanocompositeLBL self-assembly0–5
(~800 kPa)
NN500Rapid self-healing (~10 s), and 98% efficiency at room temperature.[180]
Ultrathin graphene filmSingle-step Marangoni self-assembly21037NNOptimal thickness is 4.4 nm with 3.4% failure strain.[182]
Chitosan-grapheneBath sonication and vacuum filtration0–218.6NNChitosan-G had better graphene electrical properties than pullulan and alginate.[183]

3.4. Multifunctional Sensor

Table 7 summarizes 23 carbon-based multifunctional sensors with potential applications in packaging food and medical or pharmaceutical products. Among the studies reviewed, 9 articles examined dual-functional sensors, detecting two stimuli simultaneously [185,186,187,188,189,190,191,192,193], 11 explored triple-functional sensors [194,195,196,197,198,199,200,201,202,203,204], and 3 quad-functional sensors, focusing on humidity, temperature, pressure, and strain stimuli [205,206,207]. Graphene, rGO, and CNT-based material were predominantly used, often in combination with different polymers or other materials such as carbon black to enhance the multifunctionality of the sensors. Fabrication methods often involve coating, CVD, molding, and printing, allowing for simple and scalable production. Many sensors incorporated multiple fabrication techniques to optimize performance.
The selective reactivity to specific stimuli was achieved by using different carbon materials or compositions [187,188,190,191,194,203]. For example, Bae et al. (2018) fabricated a dual-mode sensor that utilized SWCNTs/PDMS for pressure sensing (0–25 kPa) and rGO for temperature sensing (22–70 °C) with good sensitivity (0.7/kPa and 0.83%/°C, respectively), quick response (0.05 s and 0.1, respectively), and stability over 10,000 cycles [187]. Similarly, all carbon-based sensors with carbon nanocoils (CNCs) and CNTs for simultaneously sensing temperature, humidity, and strains have been developed by Li et al. [203]. The authors achieved a wide detection range from −266.15 to 126.85 °C (temperature), 10% to 80% (relative humidity) and up to 100% strain with high strain resolution (0.01%) and fast response time (16 ms) alongside a stability of 10,000 cycles. Li et al. [192] developed a dual-mode temperature and strain sensor based on graphene/PEDOT:PSS hydrogel, with a detection range of 7–60 °C for temperature and up to 1000% for strain. The sensor demonstrated high sensitivity (gauge factor 8.1 for strain, −7.16%/°C for temperature), fast response (0.2 s), and stability over 10,000 cycles.
Other reported multiple sensing systems showed a temperature range starting from 20 °C [187,188,189,190,193,197,198,202,204,205,206] or lower pressure range below 0.6 kPa [195,201], limiting their application for packaging when considering the stresses encountered in the transportation/distribution chain.
Table 7. Summary of different multifunctional sensors reported in the literature with their modes, fabrication methods, and resulting properties including the sensing range, sensitivity, durability, response, and recovery time.
Table 7. Summary of different multifunctional sensors reported in the literature with their modes, fabrication methods, and resulting properties including the sensing range, sensitivity, durability, response, and recovery time.
No. ModesCarbon MaterialsModesFabricationMechanismWorking RangeSensitivityResponse/
Recovery Time (s)
Durability (Cycles)Ref
2Monolayer grapheneHumidityCVD and oxygen plasma etchingCapacitive2–90%RH17–32%/%RH~8/~191000[185]
Resistive
TemperatureCurrent10–90 °CN~4/~10N
Cracked paddy-shaped MoS2/graphene foam/EcoflexStrainThermal CVD, dipping, and annealingPiezoresistive0–22%GF = 24.1NN[186]
Pressure0.6–25.4 kPa3.28–6.06/kPaN4000
SWCNTs/PDMSPressureCoating and moldingCapacitance0–25 kPa0.7/kPa0.05/N10,000[187]
rGOTemperatureSpray-coatingResistive22–70 °C0.83%/°C0.1/NN
PDMS/SWCNT compositePressureSpray coating and leather moldPiezoresistive0–400 kPa0.03–7.76/kPa0.132/0.1210,000 [188]
PDMS/SWCNT composite with thermochromic materialTemperatureResistive23–90 °CNNN
Graphene–CNT–Silicone adhesive nanocompositeHumidityDoctor blade and drop castingImpedance36–94% RH−84.5 Ω/%RH26/74N[189]
Capacitance1336.7 pF/%RH
TemperatureImpedance37–87 °C−19.8 Ω/°C34/82N
PU@CNT compositeTemperatureHot pressingResistive30–110 °C−2.84 × 10−3/°CN5000 [190]
PU dielectricPressureCapacitive0.1–50 kPa0.0549/kPa0.094/0.1345000
GO/SWCNTs/PDMS compositeHumidityScreen and inkjet printing
Blading and doctor blade
Resistive25–80%RH0.137–11.145%/%RH0.5/0.3N[191]
SWCNTs/PDMS compositePressurePiezoresistive0.024–230 kPa27.91–77.78 /kPa0.03/0.036000
Graphene/PEDOT:PSS hydrogelStrainOne-pot methodResistive1000%8.10.2/N10,000 [192]
Temperature7–60 °C−7.16–−0.162%/°CNN
2CNTs sponge/PEDOT:PSS/PDMSPressureSoaking and oven dryingPiezoresistive0–40 kPa26.8–902.2/kPa0.063/0.071500 [193]
TemperatureResistive20–80 °C0.84%/°C1.1/1.55 days
3GOHumiditySpray coatingCapacitive20–90%0.0589 pF/%RHNN[194]
rGOTemperatureResistive0–100 °C−3.4 kΩ/°CNN
PDMSPressureLaminationResistive0–450 kPa0.002/kPa0.2/N2000
Capacitive
rGO/CNCsCompression StrainMixing, freezing, freeze-drying, and carbonizationResistive0–99%GF = 369.4N10,000 [195]
Pressure0.00075 kPaNNN
Bending0.052–180°NN10,000
Graphene-glycerolStrainCoatingPiezoresistive0–1000%GF = 45.130.2/0.210,000 [196]
PressureResistive0–50 kPa80%NN
Twisting0–180°100%NN
PDMS-coated microporous polypyrrole/graphene foam (PDMS/PPy/GF)PressureCVD, electrochemical deposition, and dip-coatingPiezoresistive0–50 kPa2.01/kPa0.02/N10,000 [197]
TemperatureThermoelectric25–70 °C49.8 µV/K1.5/8.3N
StrainResistive0–50%GF = −1.38 (<10%)1/2.5N
GF = −0.09 (10–50%)
Carbon fibers and MWCNTs (CFs-MWCNT) compositeTemperatureResistive30–50 °C1.49–2.46%/°CNN[198]
PressurePiezoresistive0–60 kPa0.91–42.5/kPa0.1/0.16000
Bending0–180°95.5%/radN1000
GO-doped-PU nanofiber membrane coated with PEDOTPressureElectrospinning, in situ polymerization, low-temperature oxygen plasmaPiezoresistive0.001–20 kPa0.15–20.6/kPa0.012/N10,000 [199]
Strain0–550%10.1–193.2N10,000
Flexion1.0 cm−1NN6000
CNT/PDMS compositePressureReplica molding and ultraviolet-ozone exposurePiezoresistive0–270 kPa6.67/kPa0.024/0.0310,000 [200]
Bending1–6.5 mm17.7/mmNN
Tensile strain0–50%GF = 409NN
Nanopapillae-decorated carbon nanosheet (NP-CNS)HumidityPyrolysis and screen printingResistive0–96%RH8.251.7/100.1N[201]
StrainPiezoresistive0–500%GF = 21.9–99.90.07/NN
Pressure0.005–0.6 kPaN0.032/NN
3rGO/polyorganosiloxane aerogelsTemperatureCopolycondensationResistive20–100 °C50.20%N10,000 [202]
Pressure0.01–110 kPa83.50%N
Strain0.1–80%84%N
CNC (10 mg)-CNT (30 mg) buckypaperStrainMixed vacuum filtration and curingPiezoresistive0–100%GF = 352,0850.033/0.01610,000 [203]
Pre-stretched CNC (10 mg)-CNT (80 mg) buckypaperTemperatureMixed vacuum filtration, pre-stretch, and curingResistive−266.15–126.85 °C1.88%/°CN10
CNC-CNT on cellulose filter paper (1:1)HumidityDrippingResistive10–80%RHNN10
Graphene woven fabric (GWF)/PDMS compositePressureCatalytic decomposition and dippingPiezoresistive0–20 kPa0.0142/kPaN1000 [204]
Strain0–140%GF = 582NN
TemperatureThermoresistive25–80 °C0.0238/°CNN
4CB/rGO compositeStrainSpray coatingResistiveNGF = 14.6 (compression)~0.34/N1000 [205]
GF = 1.8 (tension)
Humidity16–95%RH2.04/%RH~300/~80N
Temperature20–60 °C0.6%/°C~100/N
Pressure0–250 kPa0.09–0.59%/kPa~0.25/N
CB-PDMSStrainSpin coatingResistive0–40%GF = 81.2 (0–5%)<0.05/N4000 [206]
GF = 28.5 (5–40%)
Pressure0–20 kPa4 × 104%0.1/0.1N
Flexion0–150°NN
Temperature25–150 °C0.515 ppm/°C8.4/N
Polyaniline-coated MWCNTsHumidityTwo-step assemblyConductive30–80% RH4.80%25/38
(Basal layer)
2500 [207]
56/55
(double layer)
PressurePiezoresistive0.028–100 kPaGF = 100.11/0.1310,000
Bending strain0–2.7%GF = 35.8
Twisting strain0–90°GF = 20.8

4. Discussion

There is growing research on carbon-based sensors for humidity, temperature, and mechanical stress-monitoring to improve the ability to track food and medical products during transit as illustrated by Figure 3. However, the demand for highly sensitive, durable, and scalable carbon-based sensor technologies is increasing with the global emphasis on real-time logistics monitoring. Challenges persist in translating these sensors from laboratory settings to scalable, commercially viable solutions. This systematic literature review critically assesses the recent advancements in carbon-based sensors designed to monitor humidity, temperature, and mechanical stress, either individually or as part of a multi-stimuli detection system, with potential applications in tracking food and medical products during transit. The performance and reliability of carbon-based sensors are significantly influenced by several factors, including materials selection, structural design, fabrication techniques, electrode configuration, and encapsulation strategy. These factors collectively determine the sensor’s sensitivity, responsiveness, durability, and applicability in various transportation environments. A research roadmap developed from the studies (Figure 4) shows some progress over time with efforts toward improving sensing modalities with the integration of different sensing modes as well as enhancing the functionality and properties of the sensing system, for example, with self-healing and lower power consumption or self-powered features. Alongside the need for improving the above-mentioned sensors’ performance and features, further investigation and innovations are required especially for sustainable and cost-efficient large-scale production.
The performance and reliability of carbon-based sensors is largely determined by the intrinsic properties and functional requirements of the material used. Among carbon-based materials, GO stands out for humidity sensing (Table 3) due to its low cost, large surface area and high hydrophilicity due to oxygen-containing functional groups which enhance water molecule adsorption capacity and sensitivity. However, excessive oxygen-containing groups can hinder recovery times and compromise long-term stability under high-humidity conditions [208]. Furthermore, drawbacks include potential long-term drift and low selectivity as carbon-based sensors may respond to other gases or contaminants, affecting their specificity toward water vapor.
In contrast, graphene and rGO are better suited for temperature sensors, offering exceptional electrical conductivity, thermal responsiveness, and stability. These properties enable fast response, high sensitivity, and broad detection ranges, making them suitable for most transit conditions for food and medical products including cold chain.
Mechanical sensors use the mechanical strength and piezoresistive properties of graphene for pressure sensing [170], while strain sensors benefit from the flexibility [162], conductivity, and deformation sensitivity of CNTs [173,174,176], rGO [167,177,179], and graphene [181,182]. Nonetheless, these sensors often experience structural instability and poor adhesion to substrates. The combination of GO/MWNT resulted in sensors with 1095-day stability and fast response time (0.061 s) [97]. The combination of these two materials allowed the authors to harness their complementary strengths and help offset each material’s individual limitations. Indeed graphene-based sensors typically offer faster response times and greater sensitivity but require careful structural stabilization to maintain long-term stability. In contrast, CNT-based sensors are inherently more stable, but generally exhibit slower response times. Similarly, these carbon-based materials have been used with polymers, such as PDMS [164] or PU [165], which resulted in enhanced flexibility and durability. The evolution of carbon-based sensors has increasingly shifted toward multifunctional sensing platforms, which enable simultaneous detection of multiple stimuli (e.g., pressure, strain, temperature) or analytes (e.g., gases, ions) via distinct response mechanisms, which hold significant potential for smart packaging.
Efforts to overcome these challenges have largely focused on structural design modifications and AI-driven signal processing. Multilayered architectures help minimize signal interference by physically isolating sensing components, though controlling layer thickness and interfacial properties remains a challenge. Ratiometric sensing improves accuracy by analyzing signal ratios instead of absolute values, but its reliability depends on sensor stability and calibration [209]. AI-assisted signal processing enhances detection precision through real-time filtering, noise reduction, and pattern recognition. However, integrating AI introduces challenges such as higher power consumption and computational demands, which must be addressed for practical applications. To transition from prototypes to commercial use, challenges in scalability, durability, and manufacturing must be addressed. Collaboration across materials science, engineering, and AI experts will be key to developing robust, adaptable sensors and sensing systems for real-world deployment.
Nano-structuring carbon materials, such as laser-induced graphene (LIG), graphene flowers, core–shell architectures, and nanoporous structures, have demonstrated significant advantages in sensor applications. The choice between hierarchical (e.g., nanostructured hybrids), hybrid (multi-material composites), or single-element systems (e.g., pure graphene) depends on the sensing targets and operational environments. These materials enhance sensitivity and response time through an increased surface area for the analyte interaction, improved electron transport at defect-engineered interfaces, good mechanical stability, and plasmonic amplification of certain carbon (e.g., graphene quantum dots, porous structure, or hybrid carbon–metal systems) via strong localized surface plasmon resonance effects, thermoplasmonic effects, and charge transfer [128]. The performance enhancements fundamentally stem from the intrinsic link between nanoscale architecture and function. For example, zero-dimensional (0D) quantum dots utilize size-dependent quantum confinement and unique optical properties, making them highly attractive for fluorescence-based sensing and plasmonic enhancement. Whereas one-dimensional (1D) and three-dimensional (3D) structures (e.g., carbon nanotubes, hierarchical porous carbons) leverage strong π–π interactions, offer interconnected diffusion pathways, and enable hierarchical analyte trapping to prevent aggregation, thereby ensuring efficient mass transport. Two-dimensional (2D) structures (e.g., LIG, graphene flowers) prioritize efficient charge transfer and expose abundant edge-reactive sites (such as dangling bonds and oxygenated defects), resulting in rapid binding kinetics. However, while these nanoscale structures provide clear advantages, they also introduce critical challenges. Particles below 20 nm often exhibit aggregation due to high surface energy, which compromises active surface accessibility and surface uniformity, ultimately affecting overall sensor performance [210]. Strategies like vertical alignment (e.g., CNT forests) or 3D carbon frameworks can partially address these challenges by spatially confining nanostructures while retaining their quantum confinement effects or plasmonic properties. Additionally, surface functionalization, template-assisted synthesis, and dispersion control strategies are actively being developed to improve uniformity and long-term stability. Although nanostructured carbon materials demonstrate excellent sensitivity and tunable electronic properties, they face challenges in processing complexity and stability compared to bulk carbon materials. Bulk carbon materials tend to provide higher mechanical integrity and stability, while nano-carbon materials are preferred for high-sensitivity and multifunctional sensing systems. From an economic perspective, the cost of carbon nanomaterials varies significantly due to production complexity, scalability, purity requirements, and applications [211,212]. GO is the most economical due to scalable synthesis, while high purity materials and CVD graphene command premium prices for specialized applications. rGO and MWCNTs offer a balance between cost and performance for conductive composites [212]. However, challenges in maintaining uniformity and quality during upscaling still exist (Figure 5). Advances in manufacturing technologies are critical to reducing costs and enhancing the scalability for widespread commercial applications. GQDs remain expensive due to low yields and niche use in biomedicine. As manufacturing techniques such as liquid-phase exfoliation, inkjet printing, and low-temperature plasma processes continue to advance, the cost of scalable materials like GO and rGO has stabilized. In the long term, the costs of high-purity carbon nanomaterials (such as CVD graphene and GQDs) are expected to decline with improved synthesis routes and increased market demand.
Single-element systems (e.g., pure graphene or CNTs) generally offer simplicity and cost-effectiveness but their limited tunability, low sensitivity and selectivity, and lack of multifunctionality restrict their use. Composites provide a promising and cost-effective approach to overcoming the limitations of single-material sensors by combining carbon materials with other (nano)materials such as polymers, metals or metal oxides, and other carbon additives. Hierarchical and hybrid systems enhance sensor capabilities by increasing surface area, controlling porosity, and optimizing molecular interactions, allowing for greater sensitivity and lower detection limits, while balancing complexity and scalability remains challenging [213]. Silicon polymers, particularly PDMS, are widely used in carbon-based sensors due to their exceptional chemical and thermal stability, biocompatibility, corrosion resistance, flexibility, and ease of fabrication [214]. These properties make PDMS an excellent choice for sensor substrates and encapsulation layers, allowing sensors to sustain large mechanical deformations while maintaining the integrity of the carbon sensing layer. Additionally, its hydrophobicity and chemical inertness protect sensitive materials from environmental factors, enhancing long-term durability. PDMS’s adaptability further enables the creation of complex structures and multifunctional sensing platforms. Despite these benefits, PDMS’s non-biodegradability poses environmental concerns, particularly in single-use applications, as its crosslinked structure makes recycling challenging [214]. Its fabrication processes also require significant energy or organic solvents [215], increasing its environmental footprint. Technically, achieving uniform carbon dispersion remains a challenge as it results in inappropriate composition, potentially reducing sensor sensitivity and reliability. Future research should explore alternative biodegradable polymers, such as PLA and cellulose derivatives to address these issues.
Recent research trends have highlighted the significant potential of hybrid composites, especially those that incorporate carbon materials with biopolymers, offering a promising balance between performance and sustainability. Cellulose and its derivates, including cellulose nanofiber (CNF) and cellulose nanocrystal (CNC), are promising candidates for enhancing the performance and sustainability of flexible carbon-based sensors. Cellulose’s renewability, biodegradability, and biocompatibility align with sustainability goals and ensure their safety for applications in food and medicine. Its abundant hydroxyl groups and water insolubility enable efficient water molecule adsorption, improving sensitivity and response times of carbon-based humidity sensors [84]. In addition to its role in composite reinforcement, cellulose exhibits a strong nanoscale and microscale response to humidity, which significantly impacts their mechanical properties. For example, a decrease in Young Modulus of cellulosic films with increasing RH values has be reported with CNC films showing smaller reduction, 15.6% (from 10.9 GPa to 9.2 GPa) when compared to other films such as xylan hemicellulose that showed 32.9% reduction (from 7.6 GPa to 5.1 GPa) for a change in relative humidity between 15% and 95% [216]. At the microscale, fiber swelling weakens inter fiber bonding, increasing porosity and reducing tensile strength and elastic modulus, leading to structural instability. While these moisture-induced effects are generally seen as mechanical weaknesses, they are cleverly exploited in paper-based humidity sensors. Humidity-induced cellulose swelling modifies the conductive network of embedded nanomaterials (e.g., carbon nanotubes, graphene oxide), altering electrical resistance or capacitance through nanoparticle separation or dielectric shifts [217]. This phenomenon forms the underpinning mechanism behind the development of sustainable, flexible, and highly responsive paper-based humidity sensors. In this regard, Khan et al., [77] produced a paper cellulose fiber/graphene oxide matrix (PCFGOM) humidity sensor with an increase in response to humidity ranging from 10% to 90% at 1 kHz and 10 kHz, respectively. Although the response time (1.2 s) and recovery time (0.8 s) were relatively good, further study is required as 24 h stability was reported. On the other hand, the mechanical stability of cellulose with carbon-based materials enhances flexibility, responsiveness, and durability of composites, as demonstrated in CNC-GO [64] and CNT-MC [148]. However, cellulose’s lower elasticity and stretchability compared to synthetic polymers like PDMS limits its application in mechanical sensors. Modifying cellulose, such as nanofibrillated cellulose [75], polydopamine-coated CNC [64], and methylcellulose (MC) [148], improve its compatibility with hydrophobic carbon materials like graphene, while increasing surface area and hydrophilicity. These modification enable uniform and strong integration with conductive carbon materials via hydrogen bonding or van der Waals interactions, enhancing load transfer and resistance to mechanical deformation and fatigue [75]. Despite these advantages, challenges such as poor sensing results, slow responsiveness, and scalability [77] limited long-term stability [106], and mechanical durability in high humidity or dynamic environments need to be addressed. Future research should focus on tailoring material concentrations, composite structures, and fabrication methods to achieve optimal performance. For example, incorporating ~30 wt% GO in CNF composites has been shown to maximize sensitivity while retaining flexibility [106]. Similarly, a CNT-to-MC ratio of 2:1 enables efficient temperature sensing without compromising mechanical flexibility [148]. Overall, these materials hold great potential for large-scale deployment in packaging sensors and transit monitoring systems, and advancing scalable manufacturing techniques such additive printing could further enhance the viability of these materials for industrial applications.
Fabrication methods significantly influence sensor performance, reproducibility, and industrial scalability. High-precision methods such as CVD produce high-quality films but are costly to consider and are relatively low profit-margin products such as food packaging. It is also energy-intensive, which together limits its large-scale industrial adoption. Solution-based methods such as coating and casting offer scalable and cost-effective alternatives but struggle with uniformity and reproducibility. Emerging approaches such as printing, additive manufacturing, and laser thermoforming enable scalable production, reducing material waste and allowing geometry customization, but require improvements in precision and production efficiency, especially material viscosity limitations for printing techniques. Achieving consistency, reproducibility, and scalability for large-scale production is a persistent bottleneck. Batch-to-batch variations and resource-intensive methods hinder industrial adoption [170]. Recent progress in machine learning-assisted fabrication and computational tools such as COMSOL Multiphysics 5.5 [74] have shown potential in optimizing spray-coating techniques and 3D-printing processes have enhanced fabrication accuracy and reproducibility, paving the way for scalable, multifunctional sensors.
Future research in carbon-based sensor technology specifically designed for food and medicine or pharmaceutical product packaging in transportation conditions should concentrate on several key areas to enhance sensor performance (stability, sensitivity, and selectivity) scalability and environmental sustainability. These include the development of scalable and eco-friendly fabrication methods, particularly for integrating biodegradable polymers such as PLA and cellulose derivatives with other carbon materials. Material concentrations should be tailored, and composite structures optimized to maximize sensor capabilities and mechanical stability. Hierarchical architectures, hybrid material combinations, and optimized single-element systems could be prioritized to address trade-offs between sensitivity, stability, and manufacturability, especially for multifunctional sensors on a single platform. Additionally, incorporating computational modeling and machine learning-driven optimization can accelerate the design of high-performance sensors. Finally, research should focus on long-term sensor stability under fluctuating humidity and mechanical strain conditions to ensure reliable real-world deployment.
Potential limitations of this review include the qualitative nature of the evaluation. A thematic synthesis based on the type of sensors and production methods was preferred as the variability in the reported approaches to sensor fabrication and the resulting performance metrics could render a meta-analysis impractical.

5. Conclusions

All carbon-based sensors (ACBS) for smart packaging of food and medical/pharmaceutical products are of growing interest, especially from a sustainability point of view. These sensors not only promise to enhance the safety and efficiency of supply chains but also align with increasing regulatory demands for traceability and quality assurance during transit. This review showed that humidity sensors are mainly developed with graphene oxide, whereas graphene and carbon nanotubes are predominantly used for temperature and mechanical (strains and pressure) sensors. Their performance is usually enhanced through engineering composite materials and the selection of appropriate fabrication techniques, which also determine the structural properties of the sensors. Although some progress has been made in developing all carbon-based sensors with biodegradable polymers such as cellulose, PDMS is still largely used in the reported studies, which poses environmental concerns, particularly in single-use applications. Future efforts must prioritize the development of fully biodegradable alternatives or all carbon-based sensors without PDMS with comparable properties. Cellulose and cellulose derivatives appear to be promising materials for green and sustainable sensor development and their sensitivity to moisture should be addressed. Innovations were observed in multifunctional sensor development. However, most research focused on wearable applications with dual-modal designs prevailing. This highlights clear research gaps for extending the work on wearable applications to develop sensing systems with detection ranges that meet the requirements for critical logistics scenarios like deep-frozen vaccine transport (−80 °C to −20 °C) and highly dynamic environments while ensuring scalability. Future research should focus on optimizing composite composition and structures as well as developing scalable, environmentally friendly fabrication methods to overcome current technical and commercialization barriers.
Promising prospects emerge in four key directions: (1) Sustainable hybrid systems combining biodegradable substrates such as cellulose with bio-derived conductive polymers could enable fully compostable or biodegradable sensors while maintaining performance metrics. (2) Self-healing carbon nanocomposites may revolutionize sensor durability by autonomously repairing mechanical/electrical damage during transit. (3) Integration with emerging technologies, particularly IoT-enabled blockchain tracking, AI-driven predictive analytics, and self-powered systems using triboelectric nanomaterials could transform ACBS into active components of smart logistics networks. (4) Advanced manufacturing paradigms including machine learning-assisted optimization of composite compositions and roll-to-roll manufacturing techniques may bridge the gap between lab-scale prototypes and industrial-scale production. Simultaneously, lifecycle analysis frameworks must be developed to validate the environmental benefits of ACBS against conventional electronic sensors across entire product lifetimes for packaging.
It is anticipated that cross-sector collaboration in material science, green chemistry and engineering, physics, and supply chain digitization would contribute to the next generation of ACBS to achieve parity with conventional electronic sensors within the next decade. This evolution demands a paradigm shift to embracing packaging as an active, intelligent component of food/pharma sustainable logistics ecosystems.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma18081862/s1. Table S1: PRISMA Checklist. Ref. [218] is cited in the Supplementary Materials.

Author Contributions

Conceptualization, S.G. and F.T.-M.; methodology, S.G. and F.T.-M.; formal analysis, S.G. and F.T.-M.; writing—original draft preparation, S.G., F.T.-M. and I.R.; writing—review and editing, S.G., F.T.-M., I.R., A.M.E., E.I. and Z.S.; supervision, F.T.-M., I.R., A.M.E., E.I. and Z.S. and project administration and funding acquisition F.T.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the faculty of Science and Engineering of the University of Wolverhampton and the Marie Skłodowska-Curie Research and Innovation Staff Exchange Program of the European Union under Grant Agreement No. 871163.

Acknowledgments

We are grateful to the University of Wolverhampton, Faculty of Science and Engineering as well for their financial support. The financial support of the European Research Agency through the EU Horizon 2020 MSCA RISE Project ReACTIVE Too, Grant Agreement No. 871163 is acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

Author Evgeni Ivanov was employed by the company NanoTech Lab Ltd. Author Zlatka Stoeva was employed by the company DZP Technologies Limited. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Nomenclature

3DThree-Dimensional
ACBSAll Carbon-Based Sensor
AgSilver
BPBlack Phosphorus
CBCarbon Black
CDsCarbon Dots
CFsCarbon Fibers
CMCCarboxymethyl Cellulose
CNCCellulose Nanocrystals
CNCsCarbon Nanocoils
CNFCarbon Nanofiber
CNHsCarbon Nanohorns
CNSCarbon Nanosheet
CNT−OHHydroxyl-functionalized Carbon Nanotubes
CNTsCarbon Nanotubes
Co3O4Cobalt (II,III) Oxide
CVDChemical Vapor Deposition
DADopamine
DLSDirect Laser-Scribed
EPDElectrophoretic Deposition
f-rGOFunctionalized Reduced Graphene Oxide
FGSFragmentized rGO sponge
GGraphene
g-C3N4Carbon nitride
GCMGraphene-coated microfiber
GcvdGraphene via Chemical Vapor Deposition
GFGauge Factor
GFGraphene Foam
GNPGraphene Nanoplatelets
GNRsGraphene Nanoribbons
GOGraphene Oxide
GPANIPolyaniline/Graphene
GplPlasma-Grown Graphene
GQDsGraphene Quantum Dots
GWFGraphene Woven Fabrics
HECHydroxyethyl Cellulose
HGOHummer’s Graphene Oxide
HRGOHoley-Reduced Graphene Oxide
HSMGHigh Strength Metallurgical Graphene
LBLangmuir-Blodgett
LBLLayer-by-Layer
LDWLaser Direct Writing
LiLithium
LiClLithium Chloride
LIGLaser-Induced Graphene
LiNbO3Lithium Niobate
MCMethyl Cellulose
MgMagnesium
MnManganese
MoS2Molybdenum Disulfide
MoTe2Molybdenum Ditelluride
MWCNTMulti-Walled Carbon Nanotubes
NNitrogen
NFCNanofibrillated Cellulose
NPNanopapillae
p-APp-aminophenol
P(VDF-TrFE)Poly(Vinylidene Fluoride-Trifluoroethylene)
PAMPolyacrylamide
PANIPolyaniline
PCFGOMPaper Cellulose Fiber/GO Matrix
PDAPolydopamine
PDMSPolydimethylsiloxane
PEDOT:PSSPoly(3,4-ethylenedioxythiophene) Polystyrene Sulfonate
PEGPolyethylene Glycol
PEIPolyethyleneimine or Polyetherimide
PHEAPoly-2-hydroxyethyl acrylate
POSPolyorganosiloxane
PPyPolypyrrole
PUPolyurethane
PVAPolyvinyl Alcohol
PVBPolyvinyl Butyral
PVDFPoly(vinylidene fluoride)
PVPPoly(vinylpyrrolidone)
QCMQuartz Crystal Microbalance
rGOReduced Graphene Oxide
SSulfur
SASodium Alginate
SAOSrAl2O4: Eu2+, Dy3+
SBSStyrene-Butadiene-Styrene
SDBSSodium Dodecylbenzenesulfonate
SDCShellac-derived Carbon
SDSSodium Dodecyl Sulfate
SnO2Tin(IV) Oxide
SWCNHsSingle-Walled Carbon Nanohorns
SWCNTSingle-Walled Carbon Nanotubes
TATannic Acid
TDI2,4’-Tolylene Diisocyanate
TEMPO2,2,6,6-Tetramethylpiperidine 1-oxyl
TFBGTilted Fiber Bragg Grating
TFGTilted Fiber Grating
TOCFsTEMPO-Oxidized Cellulose Fibers
VOCsVolatile Organic Compounds
WO3Tungsten Trioxide
ZnOZinc Oxide
ZrO2Zirconium Dioxide

References

  1. Janik-Karpinska, E.; Brancaleoni, R.; Niemcewicz, M.; Wojtas, W.; Foco, M.; Podogrocki, M.; Bijak, M. Healthcare Waste—A Serious Problem for Global Health. Healthcare 2023, 11, 242. [Google Scholar] [CrossRef] [PubMed]
  2. Lipinski, B. SDG TARGET 12.3 ON FOOD LOSS AND WASTE: 2023 PROGRESS REPORT|Champions 12.3; World Resources Institute: Washington, DC, USA, 2023; Available online: https://champions123.org/publication/sdg-target-123-2023-progress-report (accessed on 21 November 2024).
  3. NHS. NHS Clinical Waste Strategy; NHS England: London, UK, 2023; Available online: https://www.england.nhs.uk/long-read/nhs-clinical-waste-strategy/ (accessed on 21 November 2024).
  4. UKHSA. Vaccine Incident Guidance: Responding to Errors in Vaccine Storage, Handling and Administration; UK Health Security Agency: London, UK, 2022. Available online: https://assets.publishing.service.gov.uk/media/62c598188fa8f54e855dfe17/UKHSA-vaccine-incident-guidance-6-july-2022.pdf (accessed on 21 November 2024).
  5. WRAP. Returning to Normality After COVID-19: Food Waste Attitudes and Behaviours in 2021; WRAP: Banbury, UK, 2021; Available online: https://www.wrap.ngo/sites/default/files/2021-08/food-trends-report-august-2021.pdf (accessed on 21 November 2024).
  6. UNEP. United Nations Environment Programme Food Waste Index Report 2021; UNEP: Nairobi, Kenya, 2021; Available online: https://www.unep.org/resources/report/unep-food-waste-index-report-2021 (accessed on 21 November 2024).
  7. Food Wastage Footprint: Impacts on Natural Resources: Summary Report; FAO: Rome, Italy, 2013; ISBN 978-92-5-107752-8.
  8. Fortune Business Insights. Smart Packaging Market Size, Share & Industry Analysis, By Technology Type (Modified Atmosphere Packaging (MAP), Active Packaging, and Intelligent Packaging), By Application (Food & Beverages, Personal Care & Cosmetics, Pharmaceuticals, Electronics, E-Commerce, and Others), and Regional Forecast, 2024–2032. 2024. Available online: https://www.fortunebusinessinsights.com/smart-packaging-market-109166 (accessed on 6 November 2024).
  9. Biji, K.B.; Ravishankar, C.N.; Mohan, C.O.; Srinivasa Gopal, T.K. Smart Packaging Systems for Food Applications: A Review. J. Food Sci. Technol. 2015, 52, 6125–6135. [Google Scholar] [CrossRef] [PubMed]
  10. Ghaani, M.; Cozzolino, C.A.; Castelli, G.; Farris, S. An Overview of the Intelligent Packaging Technologies in the Food Sector. Trends Food Sci. Technol. 2016, 51, 1–11. [Google Scholar] [CrossRef]
  11. Yusufu, D.; Mills, A. A Colourimetric Vacuum Air-Pressure Indicator. Analyst 2019, 144, 5947–5952. [Google Scholar] [CrossRef]
  12. Sterr, J.; Fleckenstein, B.S.; Langowski, H.-C. The Effect of High-Pressure Processing on Tray Packages with Modified Atmosphere. Food Eng. Rev. 2015, 7, 209–221. [Google Scholar] [CrossRef]
  13. Li, X.; Koh, K.H.; Farhan, M.; Lai, K.W.C. An Ultraflexible Polyurethane Yarn-Based Wearable Strain Sensor with a Polydimethylsiloxane Infiltrated Multilayer Sheath for Smart Textiles. Nanoscale 2020, 12, 4110–4118. [Google Scholar] [CrossRef]
  14. Boz, Z.; Welt, B.A. Permeable Gas Cavity at Elevated Pressure Enhances Modified Atmosphere Packaging of Fresh Produce. J. Food Sci. 2018, 83, 1639–1649. [Google Scholar] [CrossRef]
  15. Seo, J.; Li, S.; Tsogbayar, D.; Hwang, T.; Park, J.; Ko, E.; Park, S.-J.; Yang, C.; Lee, H.S. Advanced Multiparallel-Connected Piezoresistive Physical Sensors: Elevating Performance Reliability of Flexible Strain and Pressure Sensors. ACS Appl. Mater. Interfaces 2024, 16, 22229–22237. [Google Scholar] [CrossRef]
  16. Bradford, K.J.; Dahal, P.; Van Asbrouck, J.; Kunusoth, K.; Bello, P.; Thompson, J.; Wu, F. The Dry Chain: Reducing Postharvest Losses and Improving Food Safety in Humid Climates. Trends Food Sci. Technol. 2018, 71, 84–93. [Google Scholar] [CrossRef]
  17. Kurmanov, N.; Tolysbayev, B.; Abilmazhinov, Y. The Limiting Storage Life of Perishables During Joint Transportation. CBUP 2015, 3, 499–505. [Google Scholar] [CrossRef]
  18. WHO. Good Storage and Distribution Practices; WHO: Geneva, Switzerland, 2019; Available online: https://www.who.int/publications/i/item/978-92-4-000182-4 (accessed on 13 October 2024).
  19. ISO 23412:2020; Indirect, Temperature-Controlled Refrigerated Delivery Services—Land Transport of Parcels with Intermediate Transfer. International Organization for Standardization: Geneva, Switzerland, 2020. Available online: https://www.iso.org/standard/75468.html (accessed on 14 October 2024).
  20. ISO 23416:2023; General Specifications and Testing Methods for Temperature-Sensitive Medicinal Packages in Good Distribution Practice Principles. International Organization for Standardization: Geneva, Switzerland, 2023. Available online: https://www.iso.org/obp/ui/#iso:std:iso:23416:ed-1:v1:en (accessed on 14 October 2024).
  21. ISO 22982-1:2021; Packaging—Temperature-Controlled Transport Packaging for Parcels—Part 1: General Requirements and Testing Methods. International Organization for Standardization: Geneva, Switzerland, 2021. Available online: https://www.iso.org/standard/74273.html (accessed on 14 October 2024).
  22. ISO 12048:1994; Packaging—Complete, Filled Transport Packages—Compression and Stacking Tests. International Organization for Standardization: Geneva, Switzerland, 1994. Available online: https://www.iso.org/standard/21579.html (accessed on 14 October 2024).
  23. ISO 18602:2013; Packaging and the Environment—Optimization of the Packaging System. International Organization for Standardization: Geneva, Switzerland, 2013. Available online: https://www.iso.org/standard/55870.html (accessed on 14 October 2024).
  24. ISO 13355:2016; Packaging—Complete, Filled Transport Packages and Unit Loads—Vertical Random Vibration Test. International Organization for Standardization: Geneva, Switzerland, 2016. Available online: https://www.iso.org/standard/62098.html (accessed on 14 October 2024).
  25. Tarcan, R.; Todor-Boer, O.; Petrovai, I.; Leordean, C.; Astilean, S.; Botiz, I. Reduced Graphene Oxide Today. J. Mater. Chem. C 2020, 8, 1198–1224. [Google Scholar] [CrossRef]
  26. Guo, Y.; Wei, X.; Gao, S.; Yue, W.; Li, Y.; Shen, G. Recent Advances in Carbon Material-Based Multifunctional Sensors and Their Applications in Electronic Skin Systems. Adv. Funct. Mater. 2021, 31, 2104288. [Google Scholar] [CrossRef]
  27. Hou, P.-X.; Liu, C.; Cheng, H.-M. Purification of Carbon Nanotubes. Carbon 2008, 46, 2003–2025. [Google Scholar] [CrossRef]
  28. Eatemadi, A.; Daraee, H.; Karimkhanloo, H.; Kouhi, M.; Zarghami, N.; Akbarzadeh, A.; Abasi, M.; Hanifehpour, Y.; Joo, S.W. Carbon Nanotubes: Properties, Synthesis, Purification, and Medical Applications. Nanoscale Res. Lett. 2014, 9, 393. [Google Scholar] [CrossRef] [PubMed]
  29. Rao, N.; Singh, R.; Bashambu, L. Carbon-Based Nanomaterials: Synthesis and Prospective Applications. Mater. Today Proc. 2021, 44, 608–614. [Google Scholar] [CrossRef]
  30. Nasir, S.; Hussein, M.; Zainal, Z.; Yusof, N. Carbon-Based Nanomaterials/Allotropes: A Glimpse of Their Synthesis, Properties and Some Applications. Materials 2018, 11, 295. [Google Scholar] [CrossRef]
  31. Kirchner, E.-M.; Hirsch, T. Recent Developments in Carbon-Based Two-Dimensional Materials: Synthesis and Modification Aspects for Electrochemical Sensors. Microchim. Acta 2020, 187, 441. [Google Scholar] [CrossRef]
  32. Sevilla, M.; Fuertes, A.B. The Production of Carbon Materials by Hydrothermal Carbonization of Cellulose. Carbon 2009, 47, 2281–2289. [Google Scholar] [CrossRef]
  33. Simeonova, D.D.; Pollmann, K.; Bianco, A.; Lièvremont, D. Graphene Oxide and Bacteria Interactions: What Is Known and What Should We Expect? mSphere 2024, 9, e00715-23. [Google Scholar] [CrossRef]
  34. Lv, C.; Hu, C.; Luo, J.; Liu, S.; Qiao, Y.; Zhang, Z.; Song, J.; Shi, Y.; Cai, J.; Watanabe, A. Recent Advances in Graphene-Based Humidity Sensors. Nanomaterials 2019, 9, 422. [Google Scholar] [CrossRef]
  35. Zhu, S.; Kim, D.; Jeong, C. Recent Development of Mechanical Stimuli Detectable Sensors, Their Future, and Challenges: A Review. Sensors 2023, 23, 4300. [Google Scholar] [CrossRef] [PubMed]
  36. Mitura, K.; Kornacka, J.; Kopczyńska, E.; Kalisz, J.; Czerwińska, E.; Affeltowicz, M.; Kaczorowski, W.; Kolesińska, B.; Frączyk, J.; Bakalova, T.; et al. Active Carbon-Based Nanomaterials in Food Packaging. Coatings 2021, 11, 161. [Google Scholar] [CrossRef]
  37. Kamran, U.; Heo, Y.-J.; Lee, J.W.; Park, S.-J. Functionalized Carbon Materials for Electronic Devices: A Review. Micromachines 2019, 10, 234. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, Y.; Adam, M.L.; Zhao, Y.; Zheng, W.; Gao, L.; Yin, Z.; Zhao, H. Machine Learning-Enhanced Flexible Mechanical Sensing. Nano-Micro Lett. 2023, 15, 55. [Google Scholar] [CrossRef]
  39. Yang, T.; Jiang, X.; Huang, Y.; Tian, Q.; Zhang, L.; Dai, Z.; Zhu, H. Mechanical Sensors Based on Two-Dimensional Materials: Sensing Mechanisms, Structural Designs and Wearable Applications. iScience 2022, 25, 103728. [Google Scholar] [CrossRef]
  40. Tang, C.; Wang, Y.; Li, Y.; Zeng, S.; Kong, L.; Li, L.; Sun, J.; Zhu, M.; Deng, T. A Review of Graphene-Based Temperature Sensors. Microelectron. Eng. 2023, 278, 112015. [Google Scholar] [CrossRef]
  41. European Parliament; Council of the European Union. Regulation (EU) 2025/40 of the European Parliament and of the Council of 19 December 2024 on Packaging and Packaging Waste, Amending Regulation (EU) 2019/1020 and Directive (EU) 2019/904, and Repealing Directive 94/62/EC (Text with EEA Relevance). Available online: http://data.europa.eu/eli/reg/2025/40/oj (accessed on 1 March 2025).
  42. European Parliament; Council of the European Union. European Parliament and Council Directive 94/62/EC of 20 December 1994 on Packaging and Packaging Waste; Publications Office of the European Union: Luxembourg, 2018; Volume 31994L0062, Available online: http://data.europa.eu/eli/dir/1994/62/oj (accessed on 1 March 2025).
  43. Secretary of State for Environment, Food and Rural Affairs. 2024 No. 1332 The Producer Responsibility Obligations (Packaging and Packaging Waste) Regulations 2024; The Stationery Office: London, UK, 2024. Available online: https://www.legislation.gov.uk/uksi/2024/1332/made (accessed on 1 March 2025).
  44. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 Statement: An Updated Guideline for Reporting Systematic Reviews. Syst. Rev. 2021, 10, 89. [Google Scholar] [CrossRef]
  45. Rathi, K.; Pal, K. Impact of Doping on GO: Fast Response–Recovery Humidity Sensor. ACS Omega 2017, 2, 842–851. [Google Scholar] [CrossRef]
  46. Fatima, Q.; Haidry, A.A.; Yao, Z.; He, Y.; Li, Z.; Sun, L.; Xie, L. The Critical Role of Hydroxyl Groups in Water Vapor Sensing of Graphene Oxide. Nanoscale Adv. 2019, 1, 1319–1330. [Google Scholar] [CrossRef]
  47. Serban, B.-C.; Cobianu, C.; Buiu, O.; Bumbac, M.; Dumbravescu, N.; Avramescu, V.; Nicolescu, C.M.; Brezeanu, M.; Radulescu, C.; Craciun, G.; et al. Quaternary Oxidized Carbon Nanohorns—Based Nanohybrid as Sensing Coating for Room Temperature Resistive Humidity Monitoring. Coatings 2021, 11, 530. [Google Scholar] [CrossRef]
  48. Serban, B.-C.; Cobianu, C.; Buiu, O.; Bumbac, M.; Dumbravescu, N.; Avramescu, V.; Nicolescu, C.M.; Brezeanu, M.; Pachiu, C.; Craciun, G.; et al. Ternary Nanocomposites Based on Oxidized Carbon Nanohorns as Sensing Layers for Room Temperature Resistive Humidity Sensing. Materials 2021, 14, 2705. [Google Scholar] [CrossRef] [PubMed]
  49. Zhu, J.; Cao, Y.; Chen, H.; Fan, B.; Zou, X.; Cheng, J.; Zhang, C. Rapid-Response Humidity Sensors Based on Ultra-Thin Films Stacked with Single-Layer Graphene Oxide. Results Chem. 2024, 7, 101444. [Google Scholar] [CrossRef]
  50. Trigona, C.; Al-Hamry, A.; Kanoun, O.; Baglio, S. Analysis of a Hybrid Micro-Electro-Mechanical Sensor Based on Graphene Oxide/Polyvinyl Alcohol for Humidity Measurements. Sensors 2019, 19, 1720. [Google Scholar] [CrossRef] [PubMed]
  51. Bi, H.; Yin, K.; Xie, X.; Ji, J.; Wan, S.; Sun, L.; Terrones, M.; Dresselhaus, M.S. Ultrahigh Humidity Sensitivity of Graphene Oxide. Sci. Rep. 2013, 3, 2714. [Google Scholar] [CrossRef]
  52. Li, N.; Chen, X.; Chen, X.; Ding, X.; Zhao, X. Ultrahigh Humidity Sensitivity of Graphene Oxide Combined with Ag Nanoparticles. RSC Adv. 2017, 7, 45988–45996. [Google Scholar] [CrossRef]
  53. Paterakis, G.; Vaughan, E.; Gawade, D.R.; Murray, R.; Gorgolis, G.; Matsalis, S.; Anagnostopoulos, G.; Buckley, J.L.; O’Flynn, B.; Quinn, A.J.; et al. Highly Sensitive and Ultra-Responsive Humidity Sensors Based on Graphene Oxide Active Layers and High Surface Area Laser-Induced Graphene Electrodes. Nanomaterials 2022, 12, 2684. [Google Scholar] [CrossRef]
  54. Fei, X.; Huang, J.; Shi, W. Humidity Sensor Composed of Laser-Induced Graphene Electrode and Graphene Oxide for Monitoring Respiration and Skin Moisture. Sensors 2023, 23, 6784. [Google Scholar] [CrossRef]
  55. Ni, L.; Li, X.; Cai, F.; Dong, Z.; Deng, Y.; Jiang, T.; Su, Z.; Chang, H.; Zhang, Z.; Luo, Y. Printable and Flexible Humidity Sensor Based on Graphene -Oxide-Supported MoTe2 Nanosheets for Multifunctional Applications. Nanomaterials 2023, 13, 1309. [Google Scholar] [CrossRef]
  56. Priyadharshini, B.; Valsalal, P. An Improved Humidity Sensor with GO-Mn-Doped ZnO Nanocomposite and Dimensional Orchestration of Comb Electrode for Effective Bulk Manufacturing. Nanomaterials 2022, 12, 1659. [Google Scholar] [CrossRef]
  57. Ganbold, E.; Sharma, P.K.; Kim, E.-S.; Lee, D.-N.; Kim, N.-Y. Capacitive Humidity Sensor with a Rapid Response Time on a GO-Doped P(VDF-TrFE)/LiCl Composite for Noncontact Sensing Applications. Chemosensors 2023, 11, 122. [Google Scholar] [CrossRef]
  58. Sun, C.; Shi, Q.; Yazici, M.S.; Lee, C.; Liu, Y. Development of a Highly Sensitive Humidity Sensor Based on a Piezoelectric Micromachined Ultrasonic Transducer Array Functionalized with Graphene Oxide Thin Film. Sensors 2018, 18, 4352. [Google Scholar] [CrossRef] [PubMed]
  59. Yi, R.; Peng, B.; Zhao, Y.; Nie, D.; Chen, L.; Zhang, L. Quartz Crystal Microbalance Humidity Sensors Based on Structured Graphene Oxide Membranes with Magnesium Ions: Design, Mechanism and Performance. Membranes 2022, 12, 125. [Google Scholar] [CrossRef] [PubMed]
  60. Waheed, W.; Anwer, S.; Khan, M.U.; Sajjad, M.; Alazzam, A. 2D Ti3C2Tx-MXene Nanosheets and Graphene Oxide Based Highly Sensitive Humidity Sensor for Wearable and Flexible Electronics. Chem. Eng. J. 2024, 480, 147981. [Google Scholar] [CrossRef]
  61. Zeng, S.; Pan, Q.; Huang, Z.; Gu, C.; Wang, T.; Xu, J.; Yan, Z.; Zhao, F.; Li, P.; Tu, Y.; et al. Ultrafast Response of Self-Powered Humidity Sensor of Flexible Graphene Oxide Film. Mater. Des. 2023, 226, 111683. [Google Scholar] [CrossRef]
  62. Yao, X.; Chen, L.; Luo, Z.; Ye, C.; Liang, F.; Yang, T.; Liu, X.; Tian, X.; Bi, H.; Wang, C.; et al. High-performance Flexible Humidity Sensors for Breath Detection and Non-touch Switches. Nano Sel. 2022, 3, 1168–1177. [Google Scholar] [CrossRef]
  63. Chen, Q.; Huang, X.; Yao, Y.; Mao, K. Analysis of the Effect of Electrode Materials on the Sensitivity of Quartz Crystal Microbalance. Nanomaterials 2022, 12, 975. [Google Scholar] [CrossRef]
  64. Yao, Y.; Huang, X.; Chen, Q.; Zhang, Z.; Ling, W. High Sensitivity and High Stability QCM Humidity Sensors Based on Polydopamine Coated Cellulose Nanocrystals/Graphene Oxide Nanocomposite. Nanomaterials 2020, 10, 2210. [Google Scholar] [CrossRef]
  65. Jiang, B.; Bi, Z.; Hao, Z.; Yuan, Q.; Feng, D.; Zhou, K.; Zhang, L.; Gan, X.; Zhao, J. Graphene Oxide-Deposited Tilted Fiber Grating for Ultrafast Humidity Sensing and Human Breath Monitoring. Sens. Actuators B Chem. 2019, 293, 336–341. [Google Scholar] [CrossRef]
  66. Syuhada, A.; Shamsudin, M.S.; Daud, S.; Krishnan, G.; Harun, S.W.; Aziz, M.S.A. Single-Mode Modified Tapered Fiber Structure Functionalized With GO-PVA Composite Layer for Relative Humidity Sensing. Photonic Sens. 2021, 11, 314–324. [Google Scholar] [CrossRef]
  67. Huang, X.; Leng, T.; Georgiou, T.; Abraham, J.; Raveendran Nair, R.; Novoselov, K.S.; Hu, Z. Graphene Oxide Dielectric Permittivity at GHz and Its Applications for Wireless Humidity Sensing. Sci. Rep. 2018, 8, 43. [Google Scholar] [CrossRef]
  68. Wang, C.; Jiao, C.; Wang, M.; Pan, J.; Wang, Q. GO/CNT−OH/Nafion Nanocomposite Humidity Sensor Based on the LC Wireless Method. Nanomaterials 2023, 13, 1925. [Google Scholar] [CrossRef] [PubMed]
  69. Alrammouz, R.; Podlecki, J.; Vena, A.; Garcia, R.; Abboud, P.; Habchi, R.; Sorli, B. Highly Porous and Flexible Capacitive Humidity Sensor Based on Self-Assembled Graphene Oxide Sheets on a Paper Substrate. Sens. Actuators B Chem. 2019, 298, 126892. [Google Scholar] [CrossRef]
  70. Jung, S.I.; Jang, I.R.; Ryu, C.; Park, J.; Padhan, A.M.; Kim, H.J. Graphene Oxide Decorated Multi-Frequency Surface Acoustic Wave Humidity Sensor for Hygienic Applications. Sci. Rep. 2023, 13, 6838. [Google Scholar] [CrossRef] [PubMed]
  71. Wei, Z.; Huang, J.; Chen, W.; Huang, Q. Fabrication and Characterization of Flexible Capacitive Humidity Sensors Based on Graphene Oxide on Porous PTFE Substrates. Sensors 2021, 21, 5118. [Google Scholar] [CrossRef]
  72. Wang, X.; Deng, Y.; Chen, X.; Jiang, P.; Cheung, Y.K.; Yu, H. An Ultrafast-Response and Flexible Humidity Sensor for Human Respiration Monitoring and Noncontact Safety Warning. Microsyst. Nanoeng. 2021, 7, 99. [Google Scholar] [CrossRef]
  73. Ouda, E.; Yousf, N.; Morsy, M.; Duraia, E.-S.M. Flexible Humidity Sensor Based on Light-Scribed Graphene Oxide. J. Mater. Sci. Mater. Electron. 2022, 33, 18241–18251. [Google Scholar] [CrossRef]
  74. Mohammedture, M.; Al Hashmi, S.; Lu, J.-Y.; Gutierrez, M.; Esawi, A.M.K.; Al Teneiji, M. Numerical Study of a Capacitive Graphene Oxide Humidity Sensor with Etched Configuration. ACS Omega 2021, 6, 29781–29787. [Google Scholar] [CrossRef]
  75. Yang, Y.; Su, G.; Li, Q.; Zhu, Z.; Liu, S.; Zhuo, B.; Li, X.; Ti, P.; Yuan, Q. Performance of the Highly Sensitive Humidity Sensor Constructed with Nanofibrillated Cellulose/Graphene Oxide/Polydimethylsiloxane Aerogel via Freeze Drying. RSC Adv. 2021, 11, 1543–1552. [Google Scholar] [CrossRef]
  76. Wang, F.; Wang, B.; Zhang, X.; Lu, M.; Zhang, Y.; Sun, C.; Peng, W. High Sensitivity Humidity Detection Based on Functional GO/MWCNTs Hybrid Nano-Materials Coated Titled Fiber Bragg Grating. Nanomaterials 2021, 11, 1134. [Google Scholar] [CrossRef]
  77. Khan, M.U.; Abbas, Y.; Abunahla, H.; Rezeq, M.; Alazzam, A.; Alamoodi, N.; Mohammad, B. Biocompatible Humidity Sensor Using Paper Cellulose Fiber/GO Matrix for Human Health and Environment Monitoring. Sens. Actuators B Chem. 2023, 393, 134188. [Google Scholar] [CrossRef]
  78. Khattak, Z.J.; Sajid, M.; Javed, M.; Zeeshan Rizvi, H.M.; Awan, F.S. Mass-Producible 2D Nanocomposite-Based Temperature-Independent All-Printed Relative Humidity Sensor. ACS Omega 2022, 7, 16605–16615. [Google Scholar] [CrossRef] [PubMed]
  79. Khan, S.A.; Saqib, M.; Rehman, M.M.; Mutee Ur Rehman, H.M.; Rahman, S.A.; Yang, Y.; Kim, S.; Kim, W.-Y. A Full-Range Flexible and Printed Humidity Sensor Based on a Solution-Processed P(VDF-TrFE)/Graphene-Flower Composite. Nanomaterials 2021, 11, 1915. [Google Scholar] [CrossRef] [PubMed]
  80. Rahman, S.A.; Khan, S.A.; Rehman, M.M.; Kim, W.-Y. Highly Sensitive and Stable Humidity Sensor Based on the Bi-Layered PVA/Graphene Flower Composite Film. Nanomaterials 2022, 12, 1026. [Google Scholar] [CrossRef]
  81. Saqib, M.; Ali Khan, S.; Mutee Ur Rehman, H.M.; Yang, Y.; Kim, S.; Rehman, M.M.; Young Kim, W. High-Performance Humidity Sensor Based on the Graphene Flower/Zinc Oxide Composite. Nanomaterials 2021, 11, 242. [Google Scholar] [CrossRef]
  82. Beniwal, A.; Ganguly, P.; Aliyana, A.K.; Khandelwal, G.; Dahiya, R. Screen-Printed Graphene-Carbon Ink Based Disposable Humidity Sensor with Wireless Communication. Sens. Actuators B Chem. 2023, 374, 132731. [Google Scholar] [CrossRef]
  83. Parthasarathy, P. Graphene/Polypyrrole/Carbon Black Nanocomposite Material Ink-Based Screen-Printed Low-Cost, Flexible Humidity Sensor. Emergent Mater. 2023, 6, 2053–2060. [Google Scholar] [CrossRef]
  84. Yoshida, A.; Wang, Y.-F.; Tachibana, S.; Hasegawa, A.; Sekine, T.; Takeda, Y.; Hong, J.; Kumaki, D.; Shiba, T.; Tokito, S. Printed, All-Carbon-Based Flexible Humidity Sensor Using a Cellulose Nanofiber/Graphene Nanoplatelet Composite. Carbon Trends 2022, 7, 100166. [Google Scholar] [CrossRef]
  85. Lim, W.Y.; Goh, C.-H.; Yap, K.Z.; Ramakrishnan, N. One-Step Fabrication of Paper-Based Inkjet-Printed Graphene for Breath Monitor Sensors. Biosensors 2023, 13, 209. [Google Scholar] [CrossRef]
  86. Wang, H.; Tang, C.; Xu, J. A Highly Sensitive Flexible Humidity Sensor Based on Conductive Tape and a Carboxymethyl Cellulose@graphene Composite. RSC Adv. 2023, 13, 27746–27755. [Google Scholar] [CrossRef]
  87. Andrić, S.; Tomašević-Ilić, T.; Bošković, M.V.; Sarajlić, M.; Vasiljević-Radović, D.; Smiljanić, M.M.; Spasenović, M. Ultrafast Humidity Sensor Based on Liquid Phase Exfoliated Graphene. Nanotechnology 2021, 32, 025505. [Google Scholar] [CrossRef]
  88. Lin, W.-D.; Lin, Y.-C.; Wu, R.-J.; Chavali, M. Gr/3D–ZnO Nanocomposites as Humidity Sensors with Enhanced Sensing Response. Polymers 2021, 13, 1623. [Google Scholar] [CrossRef] [PubMed]
  89. Hernández-Rivera, D.; Rodríguez-Roldán, G.; Mora-Martínez, R.; Suaste-Gómez, E. A Capacitive Humidity Sensor Based on an Electrospun PVDF/Graphene Membrane. Sensors 2017, 17, 1009. [Google Scholar] [CrossRef] [PubMed]
  90. Phan, D.-T.; Park, I.; Park, A.-R.; Park, C.-M.; Jeon, K.-J. Black P/Graphene Hybrid: A Fast Response Humidity Sensor with Good Reversibility and Stability. Sci. Rep. 2017, 7, 10561. [Google Scholar] [CrossRef] [PubMed]
  91. Lee, J.-U.; Ma, Y.-W.; Jeong, S.-Y.; Shin, B.-S. Direct Fabrication of Ultra-Sensitive Humidity Sensor Based on Hair-Like Laser-Induced Graphene Patterns. Micromachines 2020, 11, 476. [Google Scholar] [CrossRef]
  92. Lin, W.-D.; Chang, T.-C.; Wu, R.-J. Humidity Sensors Based on High Performance Graphene/Zirconium Dioxide Nanocomposite Material. Sens. Mater. 2018, 30, 1297. [Google Scholar] [CrossRef]
  93. Yu, Y.; Zhang, Y.; Jin, L.; Chen, Z.; Li, Y.; Li, Q.; Cao, M.; Che, Y.; Yang, J.; Yao, J. A Fast Response−Recovery 3D Graphene Foam Humidity Sensor for User Interaction. Sensors 2018, 18, 4337. [Google Scholar] [CrossRef]
  94. Muñoz, R.; León-Boigues, L.; López-Elvira, E.; Munuera, C.; Vázquez, L.; Mompeán, F.; Martín-Gago, J.Á.; Palacio, I.; García-Hernández, M. Acrylates Polymerization on Covalent Plasma-Assisted Functionalized Graphene: A Route to Synthesize Hybrid Functional Materials. ACS Appl. Mater. Interfaces 2023, 15, 46171–46180. [Google Scholar] [CrossRef]
  95. Shilpa, M.P.; Chethan, B.; Shetty, S.J.; Murari, M.S.; Waikar, M.R.; Sonkawade, R.G.; Gurumurthy, S.C. Highly Responsive Reduced Graphene Oxide Embedded PVDF Flexible Film-Based Room Temperature Operable Humidity Sensor. Sens. Actuators A Phys. 2024, 366, 115011. [Google Scholar] [CrossRef]
  96. Yang, H.; Ye, Q.; Zeng, R.; Zhang, J.; Yue, L.; Xu, M.; Qiu, Z.-J.; Wu, D. Stable and Fast-Response Capacitive Humidity Sensors Based on a ZnO Nanopowder/PVP-RGO Multilayer. Sensors 2017, 17, 2415. [Google Scholar] [CrossRef]
  97. Al-Hamry, A.; Lu, T.; Chen, H.; Adiraju, A.; Nasraoui, S.; Brahem, A.; Bajuk-Bogdanović, D.; Weheabby, S.; Pašti, I.A.; Kanoun, O. Ultra-Sensitive and Fast Humidity Sensors Based on Direct Laser-Scribed Graphene Oxide/Carbon Nanotubes Composites. Nanomaterials 2023, 13, 1473. [Google Scholar] [CrossRef]
  98. Papamatthaiou, S.; Argyropoulos, D.-P.; Farmakis, F.; Georgoulas, N. Investigation of the H 2 O Sensing Mechanism of DC-Operated Chemiresistors Based on Graphene Oxide and Thermally Reduced Graphene Oxide. IEEE Sens. J. 2019, 19, 7841–7848. [Google Scholar] [CrossRef]
  99. Lei, C.; Zhang, J.; Liang, T.; Liu, R.; Zhao, Z.; Xiong, J.; Yin, K. Humidity Sensor Based on rGO-SDS Composite Film. Micromachines 2022, 13, 504. [Google Scholar] [CrossRef] [PubMed]
  100. Chen, Z.; Wang, Y.; Shang, Y.; Umar, A.; Xie, P.; Qi, Q.; Zhou, G. One-Step Fabrication of Pyranine Modified- Reduced Graphene Oxide with Ultrafast and Ultrahigh Humidity Response. Sci. Rep. 2017, 7, 2713. [Google Scholar] [CrossRef]
  101. Yan, H.; Chen, Z.; Zeng, L.; Wang, Z.; Zheng, G.; Zhou, R. The Effect of rGO-Doping on the Performance of SnO2/rGO Flexible Humidity Sensor. Nanomaterials 2021, 11, 3368. [Google Scholar] [CrossRef]
  102. Kim, S.J.; Park, H.J.; Yoon, E.S.; Choi, B.G. Preparation of Reduced Graphene Oxide Sheets with Large Surface Area and Porous Structure for High-Sensitivity Humidity Sensor. Chemosensors 2023, 11, 276. [Google Scholar] [CrossRef]
  103. Du, Y.; Li, X.; Zhao, X.; Wang, N.; Li, D. Fabrication and Humidity Sensing of Reduced Graphene Oxide/Polyaniline Composite Film on Flexible Paper Substrate. Sens. Mater. 2022, 34, 2065. [Google Scholar] [CrossRef]
  104. Müller, C.; Al-Hamry, A.; Kanoun, O.; Rahaman, M.; Zahn, D.R.T.; Matsubara, E.Y.; Rosolen, J.M. Humidity Sensing Behavior of Endohedral Li-Doped and Undoped SWCNT/SDBS Composite Films. Sensors 2019, 19, 171. [Google Scholar] [CrossRef]
  105. Turkani, V.S.; Maddipatla, D.; Narakathu, B.B.; Saeed, T.S.; Obare, S.O.; Bazuin, B.J.; Atashbar, M.Z. A Highly Sensitive Printed Humidity Sensor Based on a Functionalized MWCNT/HEC Composite for Flexible Electronics Application. Nanoscale Adv. 2019, 1, 2311–2322. [Google Scholar] [CrossRef]
  106. Zhu, P.; Kuang, Y.; Wei, Y.; Li, F.; Ou, H.; Jiang, F.; Chen, G. Electrostatic Self-Assembly Enabled Flexible Paper-Based Humidity Sensor with High Sensitivity and Superior Durability. Chem. Eng. J. 2021, 404, 127105. [Google Scholar] [CrossRef]
  107. Arunachalam, S.; Gupta, A.A.; Izquierdo, R.; Nabki, F. Suspended Carbon Nanotubes for Humidity Sensing. Sensors 2018, 18, 1655. [Google Scholar] [CrossRef]
  108. Lee, Y.; Yoon, J.; Kim, Y.; Kim, D.M.; Kim, D.H.; Choi, S.-J. Humidity Effects According to the Type of Carbon Nanotubes. IEEE Access 2021, 9, 6810–6816. [Google Scholar] [CrossRef]
  109. Chaloeipote, G.; Samarnwong, J.; Traiwatcharanon, P.; Kerdcharoen, T.; Wongchoosuk, C. High-Performance Resistive Humidity Sensor Based on Ag Nanoparticles Decorated with Graphene Quantum Dots. R. Soc. Open Sci. 2021, 8, 210407. [Google Scholar] [CrossRef]
  110. Jlassi, K.; Mallick, S.; Eribi, A.; Chehimi, M.M.; Ahmad, Z.; Touati, F.; Krupa, I. Facile Preparation of N-S Co-Doped Graphene Quantum Dots (GQDs) from Graphite Waste for Efficient Humidity Sensing. Sens. Actuators B Chem. 2021, 328, 129058. [Google Scholar] [CrossRef]
  111. Morsy, M.; Gomaa, I.; Mokhtar, M.M.; ElHaes, H.; Ibrahim, M. Design and Implementation of Humidity Sensor Based on Carbon Nitride Modified with Graphene Quantum Dots. Sci. Rep. 2023, 13, 2891. [Google Scholar] [CrossRef] [PubMed]
  112. Wu, J.; Yin, C.; Zhou, J.; Li, H.; Liu, Y.; Shen, Y.; Garner, S.; Fu, Y.; Duan, H. Ultra-Thin Glass Based Flexible, Transparent and Ultra-Sensitive Surface Acoustic Wave Humidity Sensor with ZnO Nanowires and Graphene Quantum Dots. ACS Appl. Mater. Interfaces 2020, 12, 39817–39825. [Google Scholar] [CrossRef]
  113. Joshi, S.R.; Kim, B.; Kim, S.-K.; Kim, G.-H.; Song, W.; Park, K.; Shin, H. Low-Cost and Fast-Response Resistive Humidity Sensor Comprising Biopolymer-Derived Carbon Thin Film and Carbon Microelectrodes. J. Electrochem. Soc. 2020, 167, 147511. [Google Scholar] [CrossRef]
  114. Ling, T.Y.; Pu, S.H.; Fishlock, S.J.; Han, Y.; Reynolds, J.D.; McBride, J.W.; Chong, H.M.H. Sensing Performance of Nanocrystalline Graphite-Based Humidity Sensors. IEEE Sens. J. 2019, 19, 5421–5428. [Google Scholar] [CrossRef]
  115. Serban, B.C.; Buiu, O.; Dumbravescu, N.; Cobianu, C.; Avramescu, V.; Brezeanu, M.; Bumbac, M.; Nicolescu, C.M. Oxidized Carbon Nanohorns as Novel Sensing Layer for Resistive Humidity Sensor. ACSi 2020, 67, 469–475. [Google Scholar] [CrossRef]
  116. Saquib, M.; Shiraj, S.; Nayak, R.; Nirmale, A.; Selvakumar, M. Synthesis and Fabrication of Graphite/WO3 Nanocomposite-Based Screen-Printed Flexible Humidity Sensor. J. Electron. Mater. 2023, 52, 4226–4238. [Google Scholar] [CrossRef]
  117. Koskinen, T.; Juntunen, T.; Tittonen, I. Large-Area Thermal Distribution Sensor Based on Multilayer Graphene Ink. Sensors 2020, 20, 5188. [Google Scholar] [CrossRef]
  118. Wang, S.; Feng, M.; Wu, S.; Wang, Q.; Zhang, L. Highly Sensitive Temperature Sensor Based on Gain Competition Mechanism Using Graphene Coated Microfiber. IEEE Photonics J. 2018, 10, 1–8. [Google Scholar] [CrossRef]
  119. Pan, J.; Liu, S.; Zhang, H.; Lu, J. A Flexible Temperature Sensor Array with Polyaniline/Graphene–Polyvinyl Butyral Thin Film. Sensors 2019, 19, 4105. [Google Scholar] [CrossRef] [PubMed]
  120. Landi, G.; Granata, V.; Germano, R.; Pagano, S.; Barone, C. Low-Power and Eco-Friendly Temperature Sensor Based on Gelatin Nanocomposite. Nanomaterials 2022, 12, 2227. [Google Scholar] [CrossRef]
  121. Landi, G.; Pagano, S.; Granata, V.; Avallone, G.; La Notte, L.; Palma, A.L.; Sdringola, P.; Puglisi, G.; Barone, C. Regeneration and Long-Term Stability of a Low-Power Eco-Friendly Temperature Sensor Based on a Hydrogel Nanocomposite. Nanomaterials 2024, 14, 283. [Google Scholar] [CrossRef]
  122. Jasmi, F.; Azeman, N.H.; Bakar, A.A.A.; Zan, M.S.D.; Haji Badri, K.; Su’ait, M.S. Ionic Conductive Polyurethane-Graphene Nanocomposite for Performance Enhancement of Optical Fiber Bragg Grating Temperature Sensor. IEEE Access 2018, 6, 47355–47363. [Google Scholar] [CrossRef]
  123. Štulík, J.; Musil, O.; Josefík, F.; Kadlec, P. Graphene-Based Temperature Sensors–Comparison of the Temperature and Humidity Dependences. Nanomaterials 2022, 12, 1594. [Google Scholar] [CrossRef]
  124. Gong, X.; Zhang, L.; Huang, Y.; Wang, S.; Pan, G.; Li, L. Directly Writing Flexible Temperature Sensor with Graphene Nanoribbons for Disposable Healthcare Devices. RSC Adv. 2020, 10, 22222–22229. [Google Scholar] [CrossRef]
  125. Chen, X.; Zhang, R.; Wan, Z.; Wu, Z.; Song, D.; Xiao, X. Laser-Induced Graphene Based Flexible Sensing and Heating for Food Monitoring. ACS Appl. Electron. Mater. 2024, 6, 3597–3609. [Google Scholar] [CrossRef]
  126. Li, Q.; Bai, R.; Guo, L.; Gao, Y. All Laser Direct Writing Process for Temperature Sensor Based on Graphene and Silver. Front. Optoelectron. 2024, 17, 5. [Google Scholar] [CrossRef]
  127. Davaji, B.; Cho, H.D.; Malakoutian, M.; Lee, J.-K.; Panin, G.; Kang, T.W.; Lee, C.H. A Patterned Single Layer Graphene Resistance Temperature Sensor. Sci. Rep. 2017, 7, 8811. [Google Scholar] [CrossRef]
  128. Irfan, M.; Khan, Y.; Rehman, A.U.; Butt, M.A.; Khonina, S.N.; Kazanskiy, N.L. Plasmonic Refractive Index and Temperature Sensor Based on Graphene and LiNbO3. Sensors 2022, 22, 7790. [Google Scholar] [CrossRef] [PubMed]
  129. Viti, L.; Riccardi, E.; Beere, H.E.; Ritchie, D.A.; Vitiello, M.S. Real-Time Measure of the Lattice Temperature of a Semiconductor Heterostructure Laser via an On-Chip Integrated Graphene Thermometer. ACS Nano 2023, 17, 6103–6112. [Google Scholar] [CrossRef] [PubMed]
  130. Lei, S.; Su, N.; Li, M. Thermal-Resistance Effect of Graphene at High Temperatures in Nanoelectromechanical Temperature Sensors. Micromachines 2022, 13, 2078. [Google Scholar] [CrossRef] [PubMed]
  131. Han, S.; Zhou, S.; Mei, L.; Guo, M.; Zhang, H.; Li, Q.; Zhang, S.; Niu, Y.; Zhuang, Y.; Geng, W.; et al. Nanoelectromechanical Temperature Sensor Based on Piezoresistive Properties of Suspended Graphene Film. Nanomaterials 2023, 13, 1103. [Google Scholar] [CrossRef]
  132. Wang, J.; Wang, Y.; Su, N.; Li, M. Improving Consistency and Performance of Graphene-Based Devices via Al Sacrificial Layer. Colloid Interface Sci. Commun. 2023, 56, 100743. [Google Scholar] [CrossRef]
  133. Pawlak, R.; Lebioda, M.; Rymaszewski, J.; Szymanski, W.; Kolodziejczyk, L.; Kula, P. A Fully Transparent Flexible Sensor for Cryogenic Temperatures Based on High Strength Metallurgical Graphene. Sensors 2016, 17, 51. [Google Scholar] [CrossRef]
  134. Kun, H.; Bin, L.; Orban, M.; Donghai, Q.; Hongbo, Y. Accurate Flexible Temperature Sensor Based on Laser-Induced Graphene Material. Shock Vib. 2021, 2021, 9938010. [Google Scholar] [CrossRef]
  135. Fernández Sánchez-Romate, X.X.; Del Bosque García, A.; Sánchez, M.; Ureña, A. Electrical Transport Mechanisms in Graphene Nanoplatelet Doped Polydimethylsiloxane and Application to Ultrasensitive Temperature Sensors. ACS Appl. Mater. Interfaces 2023, 15, 22377–22394. [Google Scholar] [CrossRef]
  136. Liu, G.; Tan, Q.; Kou, H.; Zhang, L.; Wang, J.; Lv, W.; Dong, H.; Xiong, J. A Flexible Temperature Sensor Based on Reduced Graphene Oxide for Robot Skin Used in Internet of Things. Sensors 2018, 18, 1400. [Google Scholar] [CrossRef]
  137. Sehrawat, P.; Abid; Islam, S.S.; Mishra, P. Reduced Graphene Oxide Based Temperature Sensor: Extraordinary Performance Governed by Lattice Dynamics Assisted Carrier Transport. Sens. Actuators B Chem. 2018, 258, 424–435. [Google Scholar] [CrossRef]
  138. Neella, N.; Gaddam, V.; Nayak, M.M.; Dinesh, N.S.; Rajanna, K. Scalable Fabrication of Highly Sensitive Flexible Temperature Sensors Based on Silver Nanoparticles Coated Reduced Graphene Oxide Nanocomposite Thin Films. Sens. Actuators A Phys. 2017, 268, 173–182. [Google Scholar] [CrossRef]
  139. Barmpakos, D.; Belessi, V.; Schelwald, R.; Kaltsas, G. Evaluation of Inkjet-Printed Reduced and Functionalized Water-Dispersible Graphene Oxide and Graphene on Polymer Substrate—Application to Printed Temperature Sensors. Nanomaterials 2021, 11, 2025. [Google Scholar] [CrossRef] [PubMed]
  140. Khan, J.; Mariatti, M.; Zubir, S.A.; Rusli, A.; Manaf, A.A.; Khirotdin, R.K. Eco-Friendly Alkali Lignin-Assisted Water-Based Graphene Oxide Ink and Its Application as a Resistive Temperature Sensor. Nanotechnology 2024, 35, 055301. [Google Scholar] [CrossRef] [PubMed]
  141. Niu, Y.; Han, Y.; Cheng, H.; Xiong, Z.; Luo, B.; Ma, T.; Li, L.; Liu, S.; Chen, X.; Yi, C. Synthesized Silver Nanoparticles Decorated Reduced Graphene Oxide/Silver Ink for Aerosol Jet Printed Conformal Temperature Sensor with a Wide Sensing Range and Excellent Stability. J. Mater. Res. Technol. 2023, 25, 873–886. [Google Scholar] [CrossRef]
  142. Sehrawat, P.; Abid, A.; Islam, S.S. An Ultrafast Quantum Thermometer from Graphene Quantum Dots. Nanoscale Adv. 2019, 1, 1772–1783. [Google Scholar] [CrossRef]
  143. Liu, Q.; Tai, H.; Yuan, Z.; Zhou, Y.; Su, Y.; Jiang, Y. A High-Performances Flexible Temperature Sensor Composed of Polyethyleneimine/Reduced Graphene Oxide Bilayer for Real-Time Monitoring. Adv. Mater. Technol. 2019, 4, 1800594. [Google Scholar] [CrossRef]
  144. Xie, Z.; Li, H.; Mi, H.-Y.; Feng, P.-Y.; Liu, Y.; Jing, X. Freezing-Tolerant, Widely Detectable and Ultra-Sensitive Composite Organohydrogel for Multiple Sensing Applications. J. Mater. Chem. C 2021, 9, 10127–10137. [Google Scholar] [CrossRef]
  145. Morsy, M.; Darwish, A.G.; Mokhtar, M.M.; Elbashar, Y.; Elzwawy, A. Preparation, Investigation, and Temperature Sensing Application of rGO/SnO2/Co3O4 Composite. J. Mater. Sci. Mater. Electron. 2022, 33, 25419–25433. [Google Scholar] [CrossRef]
  146. Kumar, A.; Hsieh, P.-Y.; Shaikh, M.O.; Kumar, R.K.R.; Chuang, C.-H. Flexible Temperature Sensor Utilizing MWCNT Doped PEG-PU Copolymer Nanocomposites. Micromachines 2022, 13, 197. [Google Scholar] [CrossRef]
  147. Tsai, T.-W.; Ni, I.-C.; Wu, C.-I.; Cheng, I.-C.; Chen, J.-Z. Effect of Solution Aging on Temperature Sensitivity of CNT/PEDOT:PSS. ECS J. Solid State Sci. Technol. 2023, 12, 027001. [Google Scholar] [CrossRef]
  148. Xiao, Y.; Lin, J.; Xiao, J.; Weng, M.; Zhang, W.; Zhou, P.; Luo, Z.; Chen, L. A Multi-Functional Light-Driven Actuator with an Integrated Temperature-Sensing Function Based on a Carbon Nanotube Composite. Nanoscale 2021, 13, 6259–6265. [Google Scholar] [CrossRef] [PubMed]
  149. Turkani, V.S.; Maddipatla, D.; Narakathu, B.B.; Bazuin, B.J.; Atashbar, M.Z. A Carbon Nanotube Based NTC Thermistor Using Additive Print Manufacturing Processes. Sens. Actuators A Phys. 2018, 279, 1–9. [Google Scholar] [CrossRef]
  150. Kuzubasoglu, B.A.; Sayar, E.; Bahadir, S.K. Inkjet-Printed CNT/PEDOT:PSS Temperature Sensor on a Textile Substrate for Wearable Intelligent Systems. IEEE Sens. J. 2021, 21, 13090–13097. [Google Scholar] [CrossRef]
  151. Ben-Shimon, Y.; Ya’akobovitz, A. Flexible and Bio-Compatible Temperature Sensors Based on Carbon Nanotube Composites. Measurement 2021, 172, 108889. [Google Scholar] [CrossRef]
  152. Song, H.; Zhang, Y.; Cao, J. Sensing Mechanism of an Ionization Gas Temperature Sensor Based on a Carbon Nanotube Film. RSC Adv. 2017, 7, 53265–53269. [Google Scholar] [CrossRef]
  153. Sarma, S.; Lee, J.H. Developing Efficient Thin Film Temperature Sensors Utilizing Layered Carbon Nanotube Films. Sensors 2018, 18, 3182. [Google Scholar] [CrossRef]
  154. Yang, H.; Qi, D.; Liu, Z.; Chandran, B.K.; Wang, T.; Yu, J.; Chen, X. Soft Thermal Sensor with Mechanical Adaptability. Adv. Mater. 2016, 28, 9175–9181. [Google Scholar] [CrossRef]
  155. Leng, X.; Li, W.; Luo, D.; Wang, F. Differential Structure With Graphene Oxide for Both Humidity and Temperature Sensing. IEEE Sens. J. 2017, 17, 4357–4364. [Google Scholar] [CrossRef]
  156. Cai, C.; Qin, M. High-performance Bulk Silicon Interdigital Capacitive Temperature Sensor Based on Graphene Oxide. Electron. lett. 2013, 49, 488–490. [Google Scholar] [CrossRef]
  157. Soni, M.; Bhattacharjee, M.; Ntagios, M.; Dahiya, R. Printed Temperature Sensor Based on PEDOT: PSS-Graphene Oxide Composite. IEEE Sens. J. 2020, 20, 7525–7531. [Google Scholar] [CrossRef]
  158. Zhou, C.; Tang, N.; Zhang, X.; Fang, Y.; Jiang, Y.; Zhang, H.; Duan, X. Simultaneously Optimize the Response Speed and Sensitivity of Low Dimension Conductive Polymers for Epidermal Temperature Sensing Applications. Front. Chem. 2020, 8, 194. [Google Scholar] [CrossRef] [PubMed]
  159. Zhao, Y.; Tong, R.-J.; Chen, M.-Q.; Xia, F. Fluorescence Temperature Sensor Based on GQDs Solution Encapsulated in Hollow Core Fiber. IEEE Photon. Technol. Lett. 2017, 29, 1544–1547. [Google Scholar] [CrossRef]
  160. Zhang, L.; Lyu, S.; Zhang, Q.; Wu, Y.; Melcher, C.; Chmely, S.C.; Chen, Z.; Wang, S. Dual-Emitting Film with Cellulose Nanocrystal-Assisted Carbon Dots Grafted SrAl2O4, Eu2+, Dy3+ Phosphors for Temperature Sensing. Carbohydr. Polym. 2019, 206, 767–777. [Google Scholar] [CrossRef]
  161. Zhang, Q.; Liang, X.; Bi, W.; Pang, X.; Zhao, Y. Integrated Amorphous Carbon Film Temperature Sensor with Silicon Accelerometer into MEMS Sensor. Micromachines 2024, 15, 1144. [Google Scholar] [CrossRef]
  162. Smith, A.D.; Niklaus, F.; Paussa, A.; Schröder, S.; Fischer, A.C.; Sterner, M.; Wagner, S.; Vaziri, S.; Forsberg, F.; Esseni, D.; et al. Piezoresistive Properties of Suspended Graphene Membranes under Uniaxial and Biaxial Strain in Nanoelectromechanical Pressure Sensors. ACS Nano 2016, 10, 9879–9886. [Google Scholar] [CrossRef]
  163. Wang, J.; Zhu, Z.; Qi, Y.; Li, M. A Novel Crossbeam Structure with Graphene Sensing Element for N/MEMS Mechanical Sensors. Nanomaterials 2022, 12, 2101. [Google Scholar] [CrossRef]
  164. Kou, H.; Zhang, L.; Tan, Q.; Liu, G.; Dong, H.; Zhang, W.; Xiong, J. Wireless Wide-Range Pressure Sensor Based on Graphene/PDMS Sponge for Tactile Monitoring. Sci. Rep. 2019, 9, 3916. [Google Scholar] [CrossRef]
  165. Feng, C.; Yi, Z.; Jin, X.; Seraji, S.M.; Dong, Y.; Kong, L.; Salim, N. Solvent Crystallization-Induced Porous Polyurethane/Graphene Composite Foams for Pressure Sensing. Compos. Part B Eng. 2020, 194, 108065. [Google Scholar] [CrossRef]
  166. Deng, Z.; Gao, C.; Feng, S.; Zhang, H.; Liu, Y.; Zhu, Y.; Wang, J.; Xiang, X.; Xie, H. Highly Compressible, Light-Weight and Robust Nitrogen-Doped Graphene Composite Aerogel for Sensitive Pressure Sensors. Chem. Eng. J. 2023, 471, 144790. [Google Scholar] [CrossRef]
  167. Du, B.; Chao, Y.; Yang, K.; Li, B.; Luo, R.; Zhou, S.; Li, H. Stretchable and Tough Tannic Acid-Modified Graphene Oxide/Polyvinyl Alcohol Conductive Hydrogels for Strain and Pressure Sensors. AIP Adv. 2022, 12, 095206. [Google Scholar] [CrossRef]
  168. Wang, A.; Hu, M.; Zhou, L.; Qiang, X. Self-Powered Wearable Pressure Sensors with Enhanced Piezoelectric Properties of Aligned P(VDF-TrFE)/MWCNT Composites for Monitoring Human Physiological and Muscle Motion Signs. Nanomaterials 2018, 8, 1021. [Google Scholar] [CrossRef] [PubMed]
  169. Feng, X.; Ran, Y.; Li, X.; Xu, H.; Huang, Q.; Duan, Z.; Yuan, Z.; Jiang, Y.; Tai, H. Amorphous Carbon Derived from Daily Carbon Ink for Wide Detection Range, Low-Cost, Eco-Friendly and Flexible Pressure Sensor. Mater. Chem. Phys. 2024, 321, 129489. [Google Scholar] [CrossRef]
  170. Guo, D.; Lei, X.; Chen, H.; Yi, L.; Li, Y.; Zhao, Y.; Liu, F.; Cheng, G.J. Highly Flexible and Sensitive Pressure Sensor: Fabrication of Porous PDMS/Graphene Composite via Laser Thermoforming. Adv. Sens. Res. 2024, 3, 2300165. [Google Scholar] [CrossRef]
  171. Park, S.W.; Das, P.S.; Park, J.Y. Development of Wearable and Flexible Insole Type Capacitive Pressure Sensor for Continuous Gait Signal Analysis. Org. Electron. 2018, 53, 213–220. [Google Scholar] [CrossRef]
  172. Marshall, J.E.; Zhenova, A.; Roberts, S.; Petchey, T.; Zhu, P.; Dancer, C.E.J.; McElroy, C.R.; Kendrick, E.; Goodship, V. On the Solubility and Stability of Polyvinylidene Fluoride. Polymers 2021, 13, 1354. [Google Scholar] [CrossRef]
  173. Xu, W.; Allen, M.G. Deformable Strain Sensors Based on Patterned MWCNTs/Polydimethylsiloxane Composites. J. Polym. Sci. B Polym. Phys. 2013, 51, 1505–1512. [Google Scholar] [CrossRef]
  174. Chang, W.-T.; Yang, F.-S. Extended Width in Discontinuously Connected Polymer-Free Carbon Nanotubes Grown between Electrodes. Phys. B Condens. Matter 2015, 459, 24–28. [Google Scholar] [CrossRef]
  175. Tippo, P.; Maruthappan, M.; Ručman, S.; Jumrus, N.; Kantarak, E.; Sroila, W.; Thongsuwan, W.; Wiranwetchayan, O.; Thongpan, W.; Kumpika, T.; et al. Processing Improvement of Response and Stability of Strain Sensor Based on CNT-bioplastic Composite for Estimation of Elbow Angles. J. Appl. Polym. Sci. 2024, 141, e55892. [Google Scholar] [CrossRef]
  176. Jing, X.; Ma, Z.; Antwi-Afari, M.F.; Wang, L.; Li, H.; Mi, H.-Y.; Feng, P.-Y.; Liu, Y. Synthesis and Fabrication of Supramolecular Polydimethylsiloxane-Based Nanocomposite Elastomer for Versatile and Intelligent Sensing. Ind. Eng. Chem. Res. 2021, 60, 10419–10430. [Google Scholar] [CrossRef]
  177. Zhao, S.; Guo, L.; Li, J.; Li, N.; Zhang, G.; Gao, Y.; Li, J.; Cao, D.; Wang, W.; Jin, Y.; et al. Binary Synergistic Sensitivity Strengthening of Bioinspired Hierarchical Architectures Based on Fragmentized Reduced Graphene Oxide Sponge and Silver Nanoparticles for Strain Sensors and Beyond. Small 2017, 13, 1700944. [Google Scholar] [CrossRef]
  178. Wang, S.; Xiao, P.; Liang, Y.; Zhang, J.; Huang, Y.; Wu, S.; Kuo, S.-W.; Chen, T. Network Cracks-Based Wearable Strain Sensors for Subtle and Large Strain Detection of Human Motions. J. Mater. Chem. C 2018, 6, 5140–5147. [Google Scholar] [CrossRef]
  179. Chen, C.; Chu, F.; Zhang, Y.; Ma, M.; Sun, R.; Jia, P.; Sun, J. Fabricating Flexible Strain Sensor with Direct Writing Graphene/Carbon Nanotube Aerogel. ACS Appl. Electron. Mater. 2023, 5, 1429–1436. [Google Scholar] [CrossRef]
  180. Yao, Y.; Tai, H.; Wang, D.; Jiang, Y.; Yuan, Z.; Zheng, Y. One-Pot Preparation and Applications of Self-Healing, Self-Adhesive PAA-PDMS Elastomers. J. Semicond. 2019, 40, 112602. [Google Scholar] [CrossRef]
  181. Zhang, Y.; Lin, H.; Zhang, L.; Peng, S.; Weng, Z.; Wang, J.; Wu, L.; Zheng, L. Mechanical Exfoliation Assisted with Carbon Nanospheres to Prepare a Few-Layer Graphene for Flexible Strain Sensor. Appl. Surf. Sci. 2023, 611, 155649. [Google Scholar] [CrossRef]
  182. Li, X.; Yang, T.; Yang, Y.; Zhu, J.; Li, L.; Alam, F.E.; Li, X.; Wang, K.; Cheng, H.; Lin, C.-T.; et al. Large-Area Ultrathin Graphene Films by Single-Step Marangoni Self-Assembly for Highly Sensitive Strain Sensing Application. Adv. Funct. Mater. 2016, 26, 1322–1329. [Google Scholar] [CrossRef]
  183. Kasim, N.F.A.; Idris, W.F.W.; Abdullah, A.H.; Yusoh, K.; Ismail, Z. The Preparation of Graphene Ink from the Exfoliation of Graphite in Pullulan, Chitosan and Alginate for Strain-Sensitive Paper. Int. J. Biol. Macromol. 2020, 153, 1211–1219. [Google Scholar] [CrossRef]
  184. Muth, J.T.; Vogt, D.M.; Truby, R.L.; Mengüç, Y.; Kolesky, D.B.; Wood, R.J.; Lewis, J.A. Embedded 3D Printing of Strain Sensors within Highly Stretchable Elastomers. Adv. Mater. 2014, 26, 6307–6312. [Google Scholar] [CrossRef]
  185. Shehzad, K.; Shi, T.; Qadir, A.; Wan, X.; Guo, H.; Ali, A.; Xuan, W.; Xu, H.; Gu, Z.; Peng, X.; et al. Designing an Efficient Multimode Environmental Sensor Based on Graphene–Silicon Heterojunction. Adv. Mater. Technol. 2017, 2, 1600262. [Google Scholar] [CrossRef]
  186. Kim, S.J.; Mondal, S.; Min, B.K.; Choi, C.-G. Highly Sensitive and Flexible Strain–Pressure Sensors with Cracked Paddy-Shaped MoS2/Graphene Foam/Ecoflex Hybrid Nanostructures. ACS Appl. Mater. Interfaces 2018, 10, 36377–36384. [Google Scholar] [CrossRef]
  187. Bae, G.Y.; Han, J.T.; Lee, G.; Lee, S.; Kim, S.W.; Park, S.; Kwon, J.; Jung, S.; Cho, K. Pressure/Temperature Sensing Bimodal Electronic Skin with Stimulus Discriminability and Linear Sensitivity. Adv. Mater. 2018, 30, 1803388. [Google Scholar] [CrossRef]
  188. Abodurexiti, A.; Yang, C.; Maimaitiyiming, X. High-Performance Flexible Pressure and Temperature Sensors with Complex Leather Structure. Macro Mater. Eng. 2020, 305, 2000181. [Google Scholar] [CrossRef]
  189. Chani, M.T.S.; Karimov, K.S.; Asiri, A.M. Impedimetric Humidity and Temperature Sensing Properties of the Graphene–Carbon Nanotubes–Silicone Adhesive Nanocomposite. J. Mater. Sci. Mater. Electron. 2019, 30, 6419–6429. [Google Scholar] [CrossRef]
  190. Gao, Z.; Lou, Z.; Han, W.; Shen, G. A Self-Healable Bifunctional Electronic Skin. ACS Appl. Mater. Interfaces 2020, 12, 24339–24347. [Google Scholar] [CrossRef] [PubMed]
  191. Yang, C.; Abodurexiti, A.; Maimaitiyiming, X. Flexible Humidity and Pressure Sensors Realized by Molding and Inkjet Printing Processes with Sandwich Structure. Macro Mater. Eng. 2020, 305, 2000287. [Google Scholar] [CrossRef]
  192. Li, Z.; Huang, J.; Zhou, R.; Chen, Z.; Gao, W.; He, J.; Bao, R.; Pan, C. Temperature Decoupling of a Hydrogel-Based Strain Sensor under a Dynamic Temperature Field. Adv. Mater. Technol. 2023, 8, 2300404. [Google Scholar] [CrossRef]
  193. Zhang, Z.; Zhang, H.; Zhang, Q.; Zhao, X.; Li, B.; Zang, J.; Zhao, X.; Zhang, T. A Pressure and Temperature Dual-Parameter Sensor Based on a Composite Material for Electronic Wearable Devices. Micromachines 2023, 14, 690. [Google Scholar] [CrossRef]
  194. Ho, D.H.; Sun, Q.; Kim, S.Y.; Han, J.T.; Kim, D.H.; Cho, J.H. Stretchable and Multimodal All Graphene Electronic Skin. Adv. Mater. 2016, 28, 2601–2608. [Google Scholar] [CrossRef]
  195. Zhuo, H.; Hu, Y.; Tong, X.; Chen, Z.; Zhong, L.; Lai, H.; Liu, L.; Jing, S.; Liu, Q.; Liu, C.; et al. A Supercompressible, Elastic, and Bendable Carbon Aerogel with Ultrasensitive Detection Limits for Compression Strain, Pressure, and Bending Angle. Adv. Mater. 2018, 30, 1706705. [Google Scholar] [CrossRef]
  196. Qi, Z.; Bian, H.; Yang, Y.; Nie, N.; Wang, F. Graphene/Glycerin Solution-Based Multifunctional Stretchable Strain Sensor with Ultra-High Stretchability, Stability, and Sensitivity. Nanomaterials 2019, 9, 617. [Google Scholar] [CrossRef]
  197. Park, H.; Kim, J.W.; Hong, S.Y.; Lee, G.; Kim, D.S.; Oh, J.H.; Jin, S.W.; Jeong, Y.R.; Oh, S.Y.; Yun, J.Y.; et al. Microporous Polypyrrole-Coated Graphene Foam for High-Performance Multifunctional Sensors and Flexible Supercapacitors. Adv. Funct. Mater. 2018, 28, 1707013. [Google Scholar] [CrossRef]
  198. Wang, H.; Tao, J.; Jin, K.; Wang, X.; Dong, Y. Multifunctional Pressure/Temperature/Bending Sensor Made of Carbon Fibre-Multiwall Carbon Nanotubes for Artificial Electronic Application. Compos. Part A Appl. Sci. Manuf. 2022, 154, 106796. [Google Scholar] [CrossRef]
  199. Qi, K.; He, J.; Wang, H.; Zhou, Y.; You, X.; Nan, N.; Shao, W.; Wang, L.; Ding, B.; Cui, S. A Highly Stretchable Nanofiber-Based Electronic Skin with Pressure-, Strain-, and Flexion-Sensitive Properties for Health and Motion Monitoring. ACS Appl. Mater. Interfaces 2017, 9, 42951–42960. [Google Scholar] [CrossRef] [PubMed]
  200. Sun, K.; Ko, H.; Park, H.; Seong, M.; Lee, S.; Yi, H.; Park, H.W.; Kim, T.; Pang, C.; Jeong, H.E. Hybrid Architectures of Heterogeneous Carbon Nanotube Composite Microstructures Enable Multiaxial Strain Perception with High Sensitivity and Ultrabroad Sensing Range. Small 2018, 14, 1803411. [Google Scholar] [CrossRef]
  201. Wang, Y.; Wang, H.; Wang, H.; Zhang, M.; Liang, X.; Xia, K.; Zhang, Y. Calcium Gluconate Derived Carbon Nanosheet Intrinsically Decorated with Nanopapillae for Multifunctional Printed Flexible Electronics. ACS Appl. Mater. Interfaces 2019, 11, 20272–20280. [Google Scholar] [CrossRef]
  202. Zu, G.; Kanamori, K.; Nakanishi, K.; Huang, J. Superhydrophobic Ultraflexible Triple-Network Graphene/Polyorganosiloxane Aerogels for a High-Performance Multifunctional Temperature/Strain/Pressure Sensing Array. Chem. Mater. 2019, 31, 6276–6285. [Google Scholar] [CrossRef]
  203. Li, C.; Yang, S.; Guo, Y.; Huang, H.; Chen, H.; Zuo, X.; Fan, Z.; Liang, H.; Pan, L. Flexible, Multi-Functional Sensor Based on All-Carbon Sensing Medium with Low Coupling for Ultrahigh-Performance Strain, Temperature and Humidity Sensing. Chem. Eng. J. 2021, 426, 130364. [Google Scholar] [CrossRef]
  204. Tung, T.T.; Tran, M.T.; Pereira, A.L.C.; Cordeiro, C.M.B.; Nguyen, D.D.; Tai, N.-H.; Tran, V.V.; Hsu, C.-C.; Joshi, P.; Yoshimura, M.; et al. Graphene Woven Fabric-Polydimethylsiloxane Piezoresistive Films for Smart Multi-Stimuli Responses. Colloids Surf. B Biointerfaces 2023, 221, 112940. [Google Scholar] [CrossRef]
  205. Liu, H.; Xiang, H.; Wang, Y.; Li, Z.; Qian, L.; Li, P.; Ma, Y.; Zhou, H.; Huang, W. A Flexible Multimodal Sensor That Detects Strain, Humidity, Temperature, and Pressure with Carbon Black and Reduced Graphene Oxide Hierarchical Composite on Paper. ACS Appl. Mater. Interfaces 2019, 11, 40613–40619. [Google Scholar] [CrossRef]
  206. Lee, J.H.; Heo, J.S.; Kim, Y.; Eom, J.; Jung, H.J.; Kim, J.; Kim, I.; Park, H.; Mo, H.S.; Kim, Y.; et al. A Behavior-Learned Cross-Reactive Sensor Matrix for Intelligent Skin Perception. Adv. Mater. 2020, 32, 2000969. [Google Scholar] [CrossRef]
  207. Wang, X.; Yue, O.; Liu, X.; Hou, M.; Zheng, M. A Novel Bio-Inspired Multi-Functional Collagen Aggregate Based Flexible Sensor with Multi-Layer and Internal 3D Network Structure. Chem. Eng. J. 2020, 392, 123672. [Google Scholar] [CrossRef]
  208. Zhang, X.; Zhang, G.; Wang, F.; Chi, H. Evolution of Oxygen Content of Graphene Oxide for Humidity Sensing. Molecules 2024, 29, 3741. [Google Scholar] [CrossRef] [PubMed]
  209. Ali, A.-M.B.H.; Elmasry, M.R.; Bin Jardan, Y.A.; El-Wekil, M.M. Smart Fluorometric Sensing of Metal Contaminants in Canned Foods: A Carbon Dot-Based Dual-Response System for Quantifying Aluminum and Cobalt Ions. RSC Adv. 2025, 15, 6962–6973. [Google Scholar] [CrossRef] [PubMed]
  210. Bricha, M.; El Mabrouk, K. Effect of Surfactants on the Degree of Dispersion of MWNTs in Ethanol Solvent. Colloids Surf. A Physicochem. Eng. Asp. 2019, 561, 57–69. [Google Scholar] [CrossRef]
  211. Thines, R.K.; Mubarak, N.M.; Nizamuddin, S.; Sahu, J.N.; Abdullah, E.C.; Ganesan, P. Application Potential of Carbon Nanomaterials in Water and Wastewater Treatment: A Review. J. Taiwan Inst. Chem. Eng. 2017, 72, 116–133. [Google Scholar] [CrossRef]
  212. Meyer, E.; Bede, A.; Zingwe, N.; Taziwa, R. Metal Sulphides and Their Carbon Supported Composites as Platinum-Free Counter Electrodes in Dye-Sensitized Solar Cells: A Review. Materials 2019, 12, 1980. [Google Scholar] [CrossRef]
  213. Medrano-Lopez, J.A.; Villalpando, I.; Salazar, M.I.; Torres-Torres, C. Hierarchical Nanobiosensors at the End of the SARS-CoV-2 Pandemic. Biosensors 2024, 14, 108. [Google Scholar] [CrossRef]
  214. Ariati, R.; Sales, F.; Souza, A.; Lima, R.A.; Ribeiro, J. Polydimethylsiloxane Composites Characterization and Its Applications: A Review. Polymers 2021, 13, 4258. [Google Scholar] [CrossRef]
  215. Hong, S.; Kim, H.; Qaiser, N.; Baumli, P.; Hwang, B. A Review of Recent Progress in Fabrication Methods and Applications of Polydimethylsiloxane Sponge. J. Nat. Fibers 2023, 20, 2264497. [Google Scholar] [CrossRef]
  216. Marcuello, C.; Foulon, L.; Chabbert, B.; Aguié-Béghin, V.; Molinari, M. Atomic Force Microscopy Reveals How Relative Humidity Impacts the Young’s Modulus of Lignocellulosic Polymers and Their Adhesion with Cellulose Nanocrystals at the Nanoscale. Int. J. Biol. Macromol. 2020, 147, 1064–1075. [Google Scholar] [CrossRef]
  217. Korotcenkov, G.; Simonenko, N.; Simonenko, E.; Sysoev, V.; Brinzari, V. Paper-Based Humidity Sensors as Promising Flexible Devices, State of the Art, Part 2: Humidity-Sensor Performances. Nanomaterials 2023, 13, 1381. [Google Scholar] [CrossRef]
  218. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. BMJ 2021, 372, n71. [Google Scholar] [CrossRef]
Figure 1. Global smart packaging market share 2023 (adapted from Fortune business insights, 2024) [8].
Figure 1. Global smart packaging market share 2023 (adapted from Fortune business insights, 2024) [8].
Materials 18 01862 g001
Figure 2. Flow diagram of the adapted PRISMA approach used in this study capturing screening step and results.
Figure 2. Flow diagram of the adapted PRISMA approach used in this study capturing screening step and results.
Materials 18 01862 g002
Figure 3. Schematic illustration of packaging with integrated carbon-based sensor systems capable of tracking humidity, temperature, pressure, and mechanical shocks in real time for monitoring food and medical products throughout transportation and storage. As the products transit from the manufacturer to the consumers, sensors continuously record environmental changes and transmit data to the centralized monitoring platforms. When deviations from set thresholds are detected, the system generates immediate alerts, enabling timely interventions and corrective actions.
Figure 3. Schematic illustration of packaging with integrated carbon-based sensor systems capable of tracking humidity, temperature, pressure, and mechanical shocks in real time for monitoring food and medical products throughout transportation and storage. As the products transit from the manufacturer to the consumers, sensors continuously record environmental changes and transmit data to the centralized monitoring platforms. When deviations from set thresholds are detected, the system generates immediate alerts, enabling timely interventions and corrective actions.
Materials 18 01862 g003
Figure 4. Sensors develop roadmap for smart packaging. The evolution over time is not shown to scale, and each performance and function element is not depicted as a stage of development, but rather as a feature that research is actively progressing toward. Research in the future will focus on improving sensors performance (stability, selectivity, sensitivity, reusability) and sensor integration with different sensing modalities and miniaturized size supported by IoT and AI-driven signal processing for packaging with self-healing, low energy consumption, and self-power function. Although in many publications, durability and stability are estimated by cycling tests, durability refers to the sensor’s ability to withstand physical stress, environmental conditions, and wear overtime without degrading, whereas stability is sensor’s ability to maintain consistent performance and accuracy over time and, thus, with no drift or changes in sensitivity. However, most research currently focuses on wearable applications, with dual-modal designs prevailing, as adding more sensing modes introduces challenges such as signal interference/decoupling, increased fabrication complexity, and higher costs.
Figure 4. Sensors develop roadmap for smart packaging. The evolution over time is not shown to scale, and each performance and function element is not depicted as a stage of development, but rather as a feature that research is actively progressing toward. Research in the future will focus on improving sensors performance (stability, selectivity, sensitivity, reusability) and sensor integration with different sensing modalities and miniaturized size supported by IoT and AI-driven signal processing for packaging with self-healing, low energy consumption, and self-power function. Although in many publications, durability and stability are estimated by cycling tests, durability refers to the sensor’s ability to withstand physical stress, environmental conditions, and wear overtime without degrading, whereas stability is sensor’s ability to maintain consistent performance and accuracy over time and, thus, with no drift or changes in sensitivity. However, most research currently focuses on wearable applications, with dual-modal designs prevailing, as adding more sensing modes introduces challenges such as signal interference/decoupling, increased fabrication complexity, and higher costs.
Materials 18 01862 g004
Figure 5. Advantages and disadvantages of carbon-based sensors.
Figure 5. Advantages and disadvantages of carbon-based sensors.
Materials 18 01862 g005
Table 1. Sensing range and storage conditions for the safe transportation of food and medical/pharmaceutical products.
Table 1. Sensing range and storage conditions for the safe transportation of food and medical/pharmaceutical products.
CategoryConditionTypeRange Reference
Humidity-Dried food10–65%RH[16]
Cold chainPerishable food75–95%RH[17]
AmbientPharmaceuticals <60%RH[18]
Temperature FrozenFood −40–−18 °C[19]
Medical−40–−18 °C[19]
Cold chainFood 0–4 °C[19]
Medical2–8 °C[20]
ChilledFood 4–8 °C[19]
Medical5–25 °C[21]
AmbientFood 8–40 °C[21]
Medical15–25 °C[18]
Mechanical stress CompressionPressure34–344 kPa[22]
Strain1–15%
Impact/shockPressure5–40 G[23]
Strain1–10%
Vibration Pressure3–200 Hz[24]
Strain0.1–2% over time
Table 2. Inclusion and exclusion criteria for the screening iterations.
Table 2. Inclusion and exclusion criteria for the screening iterations.
Inclusion Criteria Exclusion Criteria
Focus on carbon-based humidity, temperature, mechanical, and multifunctional sensors for food and medical or pharmaceutical smart packagingReview articles, conference proceedings, books, and inaccessible articles
Articles discussing sensor improvement with sufficient details on sensor design, fabrication methods, and performance metricsUnrelated to sensor performance or improvements
English languagePurely theoretical articles
The properties suitable for transportation in Table 1.Articles that present speculative, unvalidated, or incomplete results
Access to full text via the authors’ institutionPublication before 2013
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, S.; Radecka, I.; Eissa, A.M.; Ivanov, E.; Stoeva, Z.; Tchuenbou-Magaia, F. Recent Advances in Carbon-Based Sensors for Food and Medical Packaging Under Transit: A Focus on Humidity, Temperature, Mechanical, and Multifunctional Sensing Technologies—A Systematic Review. Materials 2025, 18, 1862. https://doi.org/10.3390/ma18081862

AMA Style

Guo S, Radecka I, Eissa AM, Ivanov E, Stoeva Z, Tchuenbou-Magaia F. Recent Advances in Carbon-Based Sensors for Food and Medical Packaging Under Transit: A Focus on Humidity, Temperature, Mechanical, and Multifunctional Sensing Technologies—A Systematic Review. Materials. 2025; 18(8):1862. https://doi.org/10.3390/ma18081862

Chicago/Turabian Style

Guo, Siting, Iza Radecka, Ahmed M. Eissa, Evgeni Ivanov, Zlatka Stoeva, and Fideline Tchuenbou-Magaia. 2025. "Recent Advances in Carbon-Based Sensors for Food and Medical Packaging Under Transit: A Focus on Humidity, Temperature, Mechanical, and Multifunctional Sensing Technologies—A Systematic Review" Materials 18, no. 8: 1862. https://doi.org/10.3390/ma18081862

APA Style

Guo, S., Radecka, I., Eissa, A. M., Ivanov, E., Stoeva, Z., & Tchuenbou-Magaia, F. (2025). Recent Advances in Carbon-Based Sensors for Food and Medical Packaging Under Transit: A Focus on Humidity, Temperature, Mechanical, and Multifunctional Sensing Technologies—A Systematic Review. Materials, 18(8), 1862. https://doi.org/10.3390/ma18081862

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

Article Metrics

Back to TopTop