Zhen
He
ab,
Jialu
Shen
ab,
Maohua
Lan
ab and
Haibin
Gu
*ab
aKey Laboratory of Leather Chemistry and Engineering of Ministry of Education, Sichuan University, Chengdu 610065, China. E-mail: guhaibinkong@126.com
bNational Engineering Laboratory for Clean Technology of Leather Manufacture, Sichuan University, Chengdu 610065, China
First published on 13th June 2024
Conductive hydrogels (CHs) with high sensitivity and multifunctional property are considered as excellent materials for wearable devices and flexible electronics. Surface synapses and internal multilayered structures are key factors for highly sensitive pressure sensors. Nevertheless, current CHs lack environmental adaptability, multifunctional perception, and instrument portability, which seriously hinders their application as sensors. Here, waste collagen fibers (buffing dust of leather), polyvinyl alcohol (PVA) and gelatin (Gel) were used as the basic framework of the hydrogel, loaded with a conductive material (silver nanoparticles (BD-CQDs@AgNPs)) and an anti-freezing moisturizer (glycerol (Gly)), resulting in a multifunctional conductive organohydrogel (BPGC-Gly). As a temperature and humidity sensor, it demonstrated an excellent temperature response range (−20–60 °C) and was capable of rapid response (2.4 s) and recovery (1.6 s) to human breathing. As a strain/pressure sensor, it allowed real-time monitoring of human movement and had a high low-pressure sensitivity (S = 4.26 kPa−1, 0–12.5 kPa). Interestingly, BPGC-Gly could also be used as a portable bioelectrode or the acquisition, monitoring and analysis of EMG/ECG signals. In this work, BPGC-Gly was assembled with wireless transmission to achieve multimodal heath detection, which opens new avenues for multi-responsive CHs, comprehensive human health monitoring and next-generation wearable electronic skin (e-skin).
Firstly, sensitivity is one of the most critical indicators of CHs. To enhance their sensitivity, current researchers have focused on engineering CHs’ surface structures by techniques such as micro-pattern processing (micro-dome, micro-pyramid arrays and others),15 3D printing,16 pattern photolithography,17 and self-wrinkling.18 For example, Meng et al. used polypyrrole (PPy) and conductive wood to prepare a micro-pattern anti-expansion hydrogel for underwater detection of stethoscope.19 This carefully designed outer surface improves the sensitivity of CHs, but the methods are costly and require specialized equipment. Nevertheless, this approach, constructing CHs with “external protrusions”, is valuable for enhancing external perception. The natural animal biomass materials have attracted the attention of researchers due to their unique structures, such as synapses, natural fibers, pores, etc.20–22 For instance, Wei et al. prepared piezoresistive sensors with oriented multilayer structures from multilevel fiber structures contained in tannery solid wastes (TSWs).23 A unique material, buffing dust (BD), has attracted our attention; it is a leather waste with a composition similar to human skin, with abundant production, collagen-containing fiber structure and low-cost characteristics.24,25
Secondly, conductivity and antimicrobial properties need to be considered when preparing FSDs. Carbon quantum dots (CQDs), as a zero-dimensional carbon material with excellent electrical conductivity, water solubility and biocompatibility, have been widely reported in the fields of medical imaging, environmental monitoring and chemical analysis.26,27 In this work, BD-CQDs were prepared from Tara tannin, a vegetable tanning agent containing many phenolic hydroxyl structures, and BD. These BD-CQDs with a large amount of amine and catechol groups have enhanced reducing ability.28 Therefore, BD-CQDs were used as a reducing agent and stabilizer to prepare BD-CQDs silver nanoparticles (BD-CQDs@AgNPs), endowing the hydrogel with good conductivity and antibacterial properties, thereby realizing its long-term use possibility. Thirdly, to address the adaptability of CHs in different environments (low temperature and dry), strategies like hydrophobic coatings,29 salts,30 and polyols have been proposed.31 Among them, glycerin (Gly) has been widely used in food, cosmetics and medical industries because of its low cost and good moisturizing properties.32 Therefore, on the one hand, the introduction of Gly can form a strong hydrogen bond with water molecules and inhibit the evaporation of water, thus conferring anti-freezing and moisturizing properties to the hydrogel.33 On the other hand, the introduction of Gly is conducive to the improvement of hygroscopicity as well as the ability of sensing humidity change of CHs.34
In this paper, highly sensitive and multifunctional BPGC-Gly organohydrogels with “surface synapses” and “internal fibers” structures were prepared by using BD as the functional raw material for hydrogels, and by using polyvinyl alcohol (PVA), gelatin (Gel), BD-CQDs@AgNPs and Gly. The obtained BPGC-Gly organohydrogel could be used as a multi-functional sensor integrating temperature, humidity, strain, pressure and bioelectric detection. It could sense temperature (−20–60 °C) and humidity (97–46.7% RH (relative humidity)) over a wide range, and quickly respond (2.4 s) and recover (1.6 s) to different frequencies of breathing. As a strain/pressure sensor, the prepared sensor could monitor the human movement status in real time. Also, it enabled the monitoring of position and weight of varying magnitudes and exhibited high pressure sensitivity (S = 4.26 kPa−1) even under low-pressure conditions (0–12.5 kPa). In addition, owing to the high sensitivity of the BPGC-Gly bioelectrode, it could accurately collect human electrophysiological signals (electromyography (EMG) and electrocardiogram (ECG)), and further obtained more comprehensive human movement health data. These natural animal biomass-based organohydrogels show great potential in electronic skin, health monitoring, and biological electrodes.
Sample | PVA (g) | Gelatin (mg) | BD (mg) | BD-CQDs@AgNPs (mL) | H2O (mL) | Gly (mL) |
---|---|---|---|---|---|---|
PVA | 0.1 | 0 | 0 | 0 | 0.9 | |
PC | 0.1 | 0 | 0 | 0.6 | 0.9 | |
B20PG | 0.1 | 20 | 20 | 0 | 0.9 | |
B20PC | 0.1 | 0 | 20 | 0.6 | 0.9 | |
B20PG10C | 0.1 | 10 | 20 | 0.6 | 0.9 | |
B20PG20C | 0.1 | 20 | 20 | 0.6 | 0.9 | |
B20PG30C | 0.1 | 30 | 20 | 0.6 | 0.9 | |
PG20C | 0.1 | 30 | 0 | 0.6 | 0.9 | |
B10PG20C | 0.1 | 30 | 10 | 0.6 | 0.9 | |
B30PG20C | 0.1 | 30 | 30 | 0.6 | 0.9 | |
BPGC-Gly3% | 0.1 | 20 | 20 | 0.6 | 0.75 | 0.05 |
BPGC-Gly7% | 0.1 | 20 | 20 | 0.6 | 0.80 | 0.10 |
BPGC-Gly10% | 0.1 | 20 | 20 | 0.6 | 0.75 | 0.15 |
BPGC-Gly13% | 0.1 | 20 | 20 | 0.6 | 0.70 | 0.20 |
Similarly, under the same pre-preparation conditions, the pre-hydrogels were mixed well with different volumes of Gly, and then the mixed system was subjected to three repeated freezing–thawing cycles to obtain the BPGC-Gly organohydrogel. The composition of the BPGC-Glyz hydrogel is shown in Table 1, where z denotes the percentage of Gly in the total volume.
Firstly, using the mechanical strength of BPGC as a reference, we optimized the Gel and BD dosages through stretching and compression experiments. As depicted in Fig. 2a, with increasing Gel content, the hydrogel's tensile strength and strain both increased, from 0.16 MPa and 200% (PVA) to 0.5 MPa and 280% (B20PG20C). The main reason is that the addition of Gel enhanced the hydrogen bonding with PVA and BD, and could prevent fracture through effective energy dissipation during tensile deformation.36 However, further increasing the Gel dosage (B20PG30C) did not significantly improve tensile strength or elongation, probably due to the saturation of hydrogen bonding sites in PVA and BD. When comparing hydrogels' toughness and conductivity (Fig. 2b), BD-CQDs@AgNPs endowed the hydrogel with electrical conductivity, with values above 0.70 S m−1 compared to 0.28 S m−1 of the pure PVA hydrogel. Meanwhile, the change in toughness was in line with the stress strength, which increased with the increase in the amount of Gel (Fig. 2b). Notably, as shown in Fig. 2c, by comparing the compressive stress changes of hydrogels at the same compressive strain (50%), it could be found that the compressive strains of PVA, B20PC, B20PG10C, B20PG20C and B20PG30C were 0.042, 0.060, 0.072, 0.085 and 0.095 MPa, respectively. In comparison, B20PG20C and B20PG30C exhibited better compressive properties, which was consistent with the results of the stress–strain experiments. In general, considering the mechanical properties, toughness, conductivity and energy saving performance of hydrogels, we chose the amount of Gel in B20PG20C for subsequent experiments.
Similarly, the amount of BD was optimized in terms of tensile and compressive properties. As shown in Fig. 2d, with the increase of the amount of BD, the fracture strength of the hydrogel gradually increased, while the breaking elongation gradually decreased. Specifically, the active groups (amino, hydroxyl, carboxyl) contained in BD could form hydrogen bonds with the side-chain groups of PVA and Gel, thereby enhancing the tensile strength of the hydrogel.37 As displayed in Fig. 2e the toughness of P20GC, B10PG20C, B20PG20C and B30PG20C was up to 0.60 MJ m−3 and above. Simultaneously, compressive stress–strain experiments were carried out on hydrogels with different amounts of BD. Under the same compressive strain (50%), both B20PG20C and B30PG20C could reach the compressive strength of 0.085 MPa (Fig. 2f). Considering the tensile strength, compressive strength, strain, and toughness of hydrogels, the B20PG20C hydrogel (referred to as the BPGC hydrogel) was chosen to further investigate the mechanical properties of the hydrogel.
The BPGC hydrogel could be stretched to twice its original length without damage, even in the twisted and knotted state, indicating that it has good mechanical strength (Fig. 2g). Meanwhile, as depicted in Fig. 2h, the recovery ability of BPGC was characterized by tensile loading–unloading experiments under different tensile strains. Obviously, BPGC almost returned to the original state after unloading, and only a small hysteresis loop was produced, indicating its excellent recovery performance. As shown in Fig. S7a and b (ESI†), the dissipated energy was strain-dependent. As BPGC's strain increased from 20% to 100% (20% per increment), the dissipated energy corresponded to 0.35, 0.93, 2.93, 5.60, and 9.88 kJ m−3, respectively, which was due to the fact that more of the dynamic hydrogen bonds in the BPGC hydrogel were broken to release more mechanical energy with the increase of external tension.36 Meanwhile, the fatigue resistance of BPGC was studied by 20 stretching cycles under 100% strain. As depicted in Fig. 2i, only the first stretching cycle curve showed a slight hysteresis loop, which was similar to that of most hydrogels, suggesting that some of the dynamic hydrogen bonds in the BPGC hydrogel were broken.36 In the subsequent 19 tensile cycle experiments, negligible irrecoverable deformation of BPGC could be found, proving its excellent fatigue resistance. Subsequently, the dissipated energy and toughness were calculated for the 1st, 5th, 10th, 15th, and 20th cycles of the cycling process. As shown in Fig. 2j, the toughness values of the 5 groups under 100% strain were close, ranging from 56.82 to 60.50 kJ m−3. Meanwhile, the dissipated energy corresponding to the 5 cycles ranged from 7.12 to 8.25 kJ m−3, accounting for 12.50–13.50% of the total work. All these results indicated that the BPGC hydrogel has superior fatigue resistance property.
Similarly, compression experiments revealed BPGC's mechanical properties. As shown in Fig. 2k, compression-unloading experiments were performed on the BPGC hydrogel at different strains (10%, 20%, 30%, 40% and 50%). Findings consistent with the tensile-unloading experiments were observed, where the hysteresis loop of BPGC increased with increasing compressive strain (dissipation energy ranged from 0.07 to 2.35 kJ m−3, Fig. S7b, ESI†). In addition, BPGC was subjected to 20 compression-unloading cycles at 50% strain (Fig. 2l). As expected, the 20-cycle curves could almost overlap and BPGC could automatically recover to the original state without any damage, indicating its excellent compressive fatigue resistance and rebound ability. At the same time, the dissipated energy and toughness of the 1st, 5th, 10th, 15th, and 20th cycles of the cycling process are shown in Fig. 2m, and the toughness values are very close to each other, around 14.00 kJ m−3. And the corresponding dissipated energy accounted for 8.08–11.50% of the total work of each cycle. In addition, Fig. 2n displays that after 20 compression cycles (50% strain), BPGC recovered to its original state without damage. Based on the above tensile and compressive results, the exceptional fatigue resistance and recovery capacity of BPGC has proven to be the basis for subsequent multifunctional sensors and flexible bioelectrodes.
The physical and chemical interactions within BPGC-Gly were further explored by using Fourier transform infrared (FTIR) spectroscopy. As illustrated in Fig. 3f, the O–H stretching vibration peak at 3340 cm−1 could be found in the FTIR spectrum of PVA, which belonged to the intramolecular or intermolecular hydrogen bond of PVA, while the stretching vibration absorption peak corresponding to the primary alcohol C–O could be observed at 1383 cm−1.28 Obviously, the major groups and structures of collagen (methylene, carboxyl and amino) could be observed in the FTIR spectra of B, G, BPG, BPGC and BPGC-Gly. Specifically, the stretching vibration peaks of N–H and O–H could be seen near 3420 cm−1, and the asymmetric and symmetric stretching vibration absorption peaks of –CH2– could be observed at 2930 and 2855 cm−1, respectively.40 The peak near 1640–1666 cm−1 corresponded to the amide I band, which was the characteristic peak of CO stretching vibration in the carboxyl group.37 The characteristic peaks of C–N stretching vibration and N–H bending vibration (amide II band) appeared at 1553 cm−1.41 The C–N bond stretching vibration and in-plane N–H bending of the protein amide bond were recorded near 1238 cm−1 (amide III band).40 Notably, the stretching vibration peaks of O–H of these five samples were significantly deviated, and a right shift of peaks was found from the PVA hydrogel (3440 cm−1) to BPGC (3420 cm−1), indicating that a number of hydrogen bonds were formed among BD, PVA and Gel.42 After Gly was added, a blue shift of the peak appeared at 3275 cm−1 for BPGC-Gly, suggesting the further enhancement of hydrogen bonding.
Thermogravimetry (TG) and derivative thermogravimetry (DTG) were used for thermal stability testing of hydrogels and some raw materials. Firstly, as indicated in Fig. 3g, the initial partial mass loss occurring between 50 and 250 °C was attributed to the significant loss of free water. Additionally, the prominent weight loss stage observed in the range of 250–450 °C was due to the decomposition of the biomass skeleton and the breaking of intermolecular hydrogen bonds.43 Finally, at the 450–700 °C stage, it is evident that the residual mass of the hydrogels increased gradually from PVA (0.16%) to BPG (7.83%), BPGC (8.16%), and BPGC-Gly (11.49%). This indicated that the introduction of BD, Gel, BD-CQDs@AgNPs, and Gly enhanced the thermal stability of the hydrogels. Furthermore, compared to the pure PVA hydrogel, the decomposition temperature of BPG was increased by 56.7 °C, which was attributed to the introduction of BD and Gel materials which enhanced the intramolecular and intermolecular hydrogen bonding in the BPG hydrogel, resulting in the formation of a dense cross-linked network (Fig. 3h). Simultaneously, due to the incorporation of BD@CQDs@AgNPs and Gly, the decomposition temperature of BPGC (372.6 °C) and BPGC-Gly (374.3 °C) was higher than that of the BPG hydrogel (356.2 °C), demonstrating that BD@CQDs@AgNPs and Gly could also form a significant number of hydrogen bonds within the BPG structure, thereby enhancing the cross-linked network and thermal stability of the hydrogel. This finding was consistent with the results obtained from SEM.
Fig. 3i presents the X-ray photoelectron spectroscopy (XPS) full spectra of PVA, BPG, BPGC and BPGC-Gly. The characteristic peak of Ag 3d appeared in the XPS spectra of BPGC and BPGC-Gly, indicating the successful introduction of BD-CQDs@AgNPs. This finding is consistent with the EDS mapping results based on SEM. Clearly, the ratio of O/C in the PVA hydrogel is 0.20, while in BPGG and BPGC-Gly, the O/C ratio increased to 0.36 and 0.38, respectively. These findings indicated that the introduction of BD, Gel, Gly, and BD-CQDs@AgNPs enhanced the presence of polar functional groups on the surface of BPGC. To demonstrate the multiple hydrogen bonding interactions between BD, Gel, Gly and PVA, we further analyzed and compared the XPS high-resolution spectra of PVA, BPG, BPGC and BPGC-Gly. The C 1s peak of the PVA sample could be deconvoluted into three peaks, the C–C–H peak (284.80 eV), the C–OH peak (286.25 eV), and the O–CO peak (288.30 eV), where the presence of the O–CO bond may be a proof for the residue of the ethyl acetate group.44 However, after the addition of BD and Gel, the high-resolution C 1s spectrum of the BPG hydrogel exhibited three characteristic peaks with higher intensities (Fig. 3j). Particularly, the C–OH peak had shifted to a higher binding energy (BE), indicating that this subpeak may be composed of C–O and C–N bonds, which was consistent with the results obtained from FTIR. Furthermore, in the spectra of BPGC and BPGC-Gly, significant shifts could be found in the C–O/C–N and O–CO peaks, indicating that the introduction of BD-CQDs@AgNPs and Gly also enhanced the hydrogen bonding interactions among the various components. As illustrated in Fig. 3k, it could be observed that the high-resolution O 1s spectrum of the PVA sample contained two single peaks corresponding to C–O (531.06 eV) and CO (533.37 eV) bonds, whereas an increase in the binding energy of the C–O bond by 0.72 eV could be found in the BPG sample.45 Additionally, the BE of the two peaks in BPGC and BPGC-Gly also experienced displacement compared to BPG. These results collectively demonstrated the successful incorporation of BD, Gel, BD-CQDs@AgNPs and Gly into the BPGC hydrogel, and the hydrogen bonding among the various components was the main interaction (mainly relying on amino, hydroxyl, and carboxyl groups).
In addition, the weight changes of BPGC and BPGC-Gly after being placed at 20 °C for 5 d were compared. As depicted in Fig. S9f (ESI†), BPGC had a residual weight of about 12% and did not have a moisturizing effect. In contrast, the weight change rate of BPGC-Gly was no more than 30%, demonstrating its long-term moisturizing properties. Next, we tested the mechanical properties of BPGC-Gly at different placement time (0–5 d). As shown in Fig. S9g (ESI†), with increasing time, the tensile strain of BPGC-Gly decreased while its tensile stress increased. Similarly, under the same strain conditions, the compressive strength of BPGC-Gly increased (Fig. S9h, ESI†). These phenomena could be attributed to the partial evaporation of free water within BPGC-Gly, leading to partial collapse of the internal 3D network, which was consistent with the observed facts.46 Furthermore, the mechanical properties of BPGC-Gly at different temperatures (−20 °C, −10 °C, 0 °C, 10 °C, and 20 °C) were compared (Fig. S9i–k, ESI†). It was found that the tensile and compressive performance of BPGC-Gly remained almost consistent at different temperatures. These experimental phenomena indicated the importance of anti-freezing and moisture retention properties of BPGC-Gly in maintaining its good tensile and compressive strength. It overcame the dependency on the environment seen in conventional conductive hydrogels and is beneficial for monitoring human movement in low-temperature environments.
To further illustrate the biocompatibility of BPGC-Gly, common CCK-8 cytotoxicity experiments were performed.43 L929 cells were cultured with different concentrations (0.1–1.0 mg mL−1) of the BPGC-Gly extract and cell viability was recorded. As shown in Fig. 4c, the survival rates of L929 cells at all concentrations exceeded 94%, proving that BPGC-Gly was non-toxic to L929 cells. Meanwhile, LIVE/DEAD cell staining was performed on the L929 cells treated with the highest concentration (1 mg mL−1) to further evaluate the cytocompatibility of BPGC-Gly. As illustrated in Fig. 4d, only a very small number of L929 cells showed red fluorescence (dead cells) and the majority of cells exhibited green fluorescence (living cells), further demonstrating the good biocompatibility of BPGC-Gly.
In addition, the stability and sensitivity of external force feedback are important indices of flexible sensor devices. As illustrated in Fig. 5b, the change in the relative resistance of BPGC-Gly at different tensile strains (0–200% with 20%) was recorded. Specifically, BPGC-Gly could last 2–4 s for each strain gradient, with a strong stable signal output. Furthermore, by calculating the gauge factors (GF) as reported in the literature,40 a curve with the equation ΔR/R0 = 0.0011ε2 + 0.7352ε could be fitted, where ε denoted the tensile strain. Here, the quadratic curve was differentiated, and the calculation formula of GF was obtained (GF = 0.0022ε + 0.7352), indicating that BPGC-Gly had strain sensitivity and offered the possibility of human motion detection (Fig. 5c). Meanwhile, the pressure sensitivity (S) was further determined to quantify the compression sensitivity of BPGC-Gly. As indicated in Fig. 5f, the absolute change in the relative resistance of BPGC-Gly under different pressure (0–125 kPa) was tested and documented. As shown in Fig. 5g, the S could be divided into three regions. A pressure sensitivity of 4.26 kPa−1 was observed in the pressure range of 0–12.5 kPa. In the moderate pressure range (12.5–55.5 kPa), the pressure sensitivity was reduced to 0.86 kPa−1. As the pressure increased to 55.5–125 kPa, the S dropped to 0.18 kPa−1. These results indicated that BPGC-Gly was highly sensitive to pressure changes in the low-pressure range. Compared to other pressure-sensitive hydrogels, it exhibited a wider range of strain detection, demonstrating its potential as a pressure sensor (Table S3, ESI†).
Furthermore, the rapid recovery capability is another crucial indicator for strain/pressure sensors. We performed stretching-release and compression-release experiments on BPGC-Gly to determine its response time. As shown in Fig. 5d and h, BPGC-Gly demonstrated swift response and recovery to both stretching and compression. Specifically, it had a response time of 750 ms and a recovery time of 300 ms during the stretching experiment. Furthermore, in the compression experiment, it maintained a difference in response and recovery time of no more than 200 ms. These results demonstrated the excellent responsiveness of BPGC-Gly to stress/pressure. Lastly, the cyclic stability and fatigue resistance of BPGC-Gly as a strain/pressure sensor were investigated. As depicted in Fig. 5e and i, the stretching/compression cycling tests were conducted at 50% strain for 500 s/2000 s. Throughout the testing process, there was very little performance loss observed, and the magnified images of the curves revealed nearly identical signals. These results suggested that BPGC-Gly possessed outstanding durability under stretching and compression, making it suitable as a flexible sensor for long-term motion detection and heavy load monitoring.
Further systematic investigations were conducted on the performance of the BPGC-Gly sensor in relative humidity detection. The detailed working principle of the BPGC-Gly humidity sensor is presented in Fig. 6e. In low relative humidity (RH) environments, the low vapor pressure induces the evaporation of some free water from BPGC-Gly, impeding the migration of internal ions and leading to the increased resistance. However, the presence of Gly in BPGC-Gly locks the moisture, preventing excessive water loss and maintaining stability in low RH environments. In high RH environments, due to the highly hydrophilic groups (amino, carboxyl, and hydroxyl groups) within BPGC-Gly, water molecules adsorb and condense on the hydrogel through chemical and physical means, causing a slight increase in the volume of the organohydrogel.49 Overall, the BPGC-Gly humidity sensor utilizes its moisture absorption and release properties to detect the RH changes. The conductance variation reflected the humidity level, allowing for accurate humidity sensing. Similarly, the RH changes could be observed through the change of electrical conductivity. As illustrated in Fig. 6f, with the increase of environmental RH, the conductivity of the BPGC-Gly humidity sensor also increased. Specifically, the conductivity increased from 0.15 S m−1 at 24.2% RH to 0.85 S m−1 at 97% RH. Furthermore, as shown in the inset of Fig. 6f, the conductivity of the BPGC-Gly humidity sensor was also effectively maintained in an open environment (21 °C, 60% RH). In addition, the conductance changes under different RH conditions were studied to evaluate the humidity sensing performance of the BPGC-Gly sensor. As illustrated in Fig. 6g, starting with an initial 24.2% RH, the BPGC-Gly sensor was placed in chambers with different RH levels (97%, 84%, 75%, 60%, and 46.7% RH) until the conductance no longer changed, and then transferred back to the initial environment until it recovered to the initial conductance value. Notably, the response ranges of the BPGC-Gly sensor were 12% and 2% in environments with 97% RH and 46.7% RH, respectively, demonstrating a wide humidity response range. Additionally, the humidity cycling stability of the BPGC-Gly sensor was also analyzed. As depicted in Fig. 6h, the standard deviation of the response curves for various humidity changes within 30 cycles did not exceed 3%, proving the reliability and stability of the humidity response of the BPGC-Gly humidity sensor. Especially, the response time and recovery time of the BPGC-Gly humidity sensor from RH 24.2% to RH 84% and then returning to the RH 24.2% environment were 104.2 s and 94.2 s, respectively, indicating that the BPGC-Gly sensor has the ability to quickly sense the humidity change (Fig. 6i).
To further explore the practical application potential of the BPGC-Gly sensor, it was fixed onto a mask to monitor real-time human respiration. As shown in Fig. 6j and k, the changes of relative conductance with time were clear and repeatable for different breathing patterns (irregular and regular modes), which proved that the BPGC-Gly sensor has the ability to recognize breath environments with different lengths and intensities. Meanwhile, due to the small size of the humidity chamber within the mask, the airflow generated by human respiration easily changes the RH around the sensor. Therefore, compared to regular chamber humidity tests, the time required for mask respiration detection was shorter. As depicted in Fig. 6k, under the regular human breathing mode, the response time of exhaled gas was 2.4 s, and the response time of inhaled gas was 1.6 s, indicating that BPGC-Gly possesses fast respiratory responsiveness. In conclusion, these results demonstrated that the BPGC-Gly sensor, with its dual sensing capabilities for temperature and humidity, has potential applications in diagnosing respiratory system diseases and sleep monitoring.
Based on the previous experiments, we know that BPGC-Gly exhibited pressure hypersensitivity, rapid responsiveness, and long-term stability. To further explore the potential of BPGC-Gly as a flexible pressure electronic device, we conducted a detailed analysis of its rapid responsiveness and stable feedback. As illustrated in Fig. 8a, under different frequencies of pressure stimulation, the peak heights of the curves of |ΔR|/R0 changing over time remained unchanged. Moreover, with increasing frequency, the response and recovery time of the signal peak also become shorter, indicating the reproducible and rapid feedback properties of BPGC-Gly at different frequencies. Additionally, during the continuous testing process under different pressures, almost no performance loss or interruption of continuous feedback was observed (Fig. 8b). BPGC-Gly with a highly porous structure exhibited a wide range of pressure (8.5–125 kPa) durability and stable feedback during compression. Its rapid responsiveness was demonstrated by compressing it from 0% to 50% and restoring it to its original state. As shown in Fig. 8c, BPGC-Gly could be almost restored to the original state and produce a fast response curve under rapid and sustained pressure. The aforementioned tests, including varying compression frequency, pressure, and compression-recovery tests, further confirm that BPGC-Gly possesses remarkable sensitivity and stability, making it a promising candidate for various applications in flexible pressure electronic devices.
BPGC-Gly pieces were coupled with an electrochemical workstation to create a 4 × 4 matrix pressure sensor for detecting pressure distribution in space. The working principle was based on the compression of BPGC-Gly under external pressure, which increases the contact area and shortens the ion flow path, resulting in a change in BPGC-Gly's conductivity.50 Each BPGC-Gly sensor was shaped as a cylindrical patch (15 × 15 × 10 mm) for individual operation. When a 50 g weight was placed on a single unit of the array sensor (Fig. 8d), the 2D mapping pressure distribution (Fig. 8e) and the 3D pressure signal distribution (Fig. 8f) could be clearly observed. Similarly, by placing different weights (100 g, 50 g, 20 g, and 10 g) on the BPGC-Gly pressure array, the response signal decreased as the weight decreased. Notably, the BPGC-Gly array sensor could also accurately identify different external intensity pressures and integrate into 2D and 3D mapping images. As shown in Fig. 8g–i, by placing different rectangular blocks of varying quantities on the pressure sensor unit, accurate pressure position distribution and relative resistance intensity could be obtained graphically. The visualization application of the BPGC-Gly pressure array sensor has potential applications in the human–computer interaction system of multi-touch devices.
Fig. 9a illustrates the mechanism of using BPGC-Gly as a flexible electrode for bioelectric detection. Let's take muscle cells as an example. It is well known that muscle contraction begins in the central nervous system, where alpha motor neurons generate muscle contraction signals (neurons are mainly composed of dendrites, cell bodies, and axons).52 As shown in the left picture of Fig. 9a, when the signal has not reached the cell body, the neuron exhibits a resting potential (approximately −70 mV due to the outflow of K+ ions from the cell membrane). The neuron is in a polarized state at this time. However, when an organism is stimulated by a stimulus, such as a nerve impulse or chemical signal, the organism generates an excitatory or inhibitory electric signal along the motor neuron. When the electrical signal reaches the cell body, the balance of the polarized state is broken, and a significant influx of Na+ from outside the cell membrane causes an increase in the membrane potential of the muscle fiber (about +40 mV), known as depolarization (in the right picture of Fig. 9a). Lastly, the potential at this time is called the action potential, and its generation causes a change in the transmembrane potential difference, which prompts the electrons in the BPGC-Gly flexible electrode attached to the skin to move directionally.37,52 Therefore, the electrochemical signal curve of the generated bioelectricity can be obtained based on the potential difference between the reference and working electrodes.
Fig. 9b depicts the hydrogel-skin tissue equivalent model for bioelectrical recording. Generally, human skin tissue cells consist of stratum corneum, epidermal cells, and dermal tissue.53 These three structures could generally be regarded as body impedance and skin impedance. The skin impedance is composed of the equivalent resistance Rs (which has a higher impedance and can reach resistivity values of 100–1000 KΩ m) of the stratum corneum and the equivalent capacitance Cs between the epidermis and dermis.54,55 Since the dermis and subcutaneous tissue mainly consist of cells and a large amount of extracellular matrix, their equivalent impedance is smaller and can be represented as the resistance Rh.37,52 Additionally, Fig. 9c demonstrates the process of bioelectrical signal acquisition, detection, and analysis. Specifically, the electrical signals generated by movement are conducted through neural and muscular tissues, causing electron migration within the BPGC-Gly flexible electrode, which is in contact with the human epidermis. Subsequently, bioelectric signals are acquired by collecting the potential difference between the working and reference electrodes. And then, these signals were conveyed to the core control circuitry through the analog-to-digital conversion circuitry, and transmitted to a computer terminal in real-time via Bluetooth for real-time data detection and analysis. Compared to traditional bioelectrodes, the BPGC-Gly electrode exhibits excellent sensitivity, flexibility, stability, reusability, and low impedance. These characteristics are advantageous for maintaining a high signal-to-noise ratio in flexible electrodes, thereby enhancing the clarity and reliability of bioelectrical signals.
As illustrated in Fig. 9d and e, the bioelectric acquisition device based on BPGC-Gly was used for muscle signal detection at room temperature (25 °C) and low temperature (−20 °C). It could be found that prolonged elbow flexion training showed stable and repetitive EMG signals, demonstrating the ability of the BPGC-Gly flexible electrode to monitor subtle muscle signals of the human body, as well as its tolerance to low-temperature environments. Based on the portability and sensitivity of the BPGC-Gly bioelectrode acquisition device, we have attempted to apply it to motion monitoring in other parts of the human body. As depicted in Fig. 9f and g, BPGC-Gly could adhere completely to the arm skin for signal transmission. We placed the reference electrode adjacent to the working electrode and grounded the ground electrode on the arm bone. By performing repetitive weightlifting and cup-grasping motions, stable motion signal curves were obtained. Through comparison, it could be observed that different motion activities of the same muscle resulted in varying peak signal intensities. This phenomenon allowed us to identify different actions on the same body part based on peak intensity. Additionally, we further utilized BPGC-Gly for monitoring large-scale body movements such as walking and running. As shown in Fig. 9h and i, the bioelectrical signal intensity and waveform obtained during fast walking and fast running (with consistent stride length) are essentially the same. However, due to the higher frequency of running movements, the corresponding electrical signals displayed a faster frequency state. Notably, the ECG signals of volunteers could also be clearly and completely detected, and the heart rate of the volunteers could be predicted by detecting the signal peak, which was compared with the heart rate of the smart bracelet. It was found that they both showed 68 beats per minute (Fig. 9j). Meanwhile, the magnified ECG could display characteristic peaks of the heartbeat, including P-wave, QRS complex and T-wave, which are key indicators of symptoms such as arrhythmia and myocarditis (Fig. 9k). Due to the importance of the T/R ratio and signal-to-noise ratio (SNR) as key indicators of accuracy in ECG detection devices, a comparison was made between the BPGC-Gly electrode (BE) and commercial electrode pads (CE) in terms of these two parameters. As shown in Fig. 9l, the ECG waveform detected by CE was consistent in shape with that of BE. Additionally, the T/R ratio for BE and CE was measured to be 0.22 and 0.24, respectively, while the SNR was reported as 10.29 and 10.13, respectively (Fig. 9m). These results indicated that BE exhibited comparable monitoring accuracy to commercial electrodes, making it suitable for ECG/EMG monitoring during muscle movements. The above findings showed that BPGC-Gly has excellent micro signal sensing performance and is expected to be used as a portable bioelectrode and for human–computer interaction, human motion monitoring, and heart disease monitoring through a complete data acquisition, reception, and analysis process. Finally, to showcase the versatility of the BPGC-Gly sensor, it was qualitatively and quantitatively compared with previously reported hydrogel sensors (Fig. 9n and Table S4, ESI†). The results indicated that the BPGC-Gly sensor possessed excellent properties, including anti-freezing, moisturizing, antibacterial, and biocompatibility. Additionally, BPGC-Gly exhibited a diverse range of multifunctional sensing capabilities, encompassing temperature, humidity, strain sensing, pressure sensing, and biological signals. This brings new possibilities to the field of multifunctional conductive hydrogels.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tb00374h |
This journal is © The Royal Society of Chemistry 2024 |