Next Article in Journal
Smartphone-Based Portable Bio-Chemical Sensors: Exploring Recent Advancements
Previous Article in Journal
Anodic Stripping Voltammetric Determination of Copper Ions in Cell Culture Media: From Transwell® to Organ-on-Chip Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress on Electrochemical Sensing of Pharmaceutical Drugs in Complex Biofluids

1
School of Chemical, Biological and Environmental Engineering, Oregon State University, Corvallis, OR 97331, USA
2
School of Electrical Engineering and Computer Science, Oregon State University, Corvallis, OR 97331, USA
*
Author to whom correspondence should be addressed.
Chemosensors 2023, 11(8), 467; https://doi.org/10.3390/chemosensors11080467
Submission received: 3 July 2023 / Revised: 14 August 2023 / Accepted: 17 August 2023 / Published: 21 August 2023
(This article belongs to the Section Electrochemical Devices and Sensors)

Abstract

:
Electrochemical detection, with its advantages of being rapid, multi-time point, compatible with cost-effective fabrication methods, and having the potential for miniaturization and portability, has great promise for point-of-care drug monitoring. However, a continuing challenge concerns the robust and sensitive electrochemical detection of pharmaceutical analytes from biological fluids. These complex matrices, such as saliva, sweat, interstitial fluid, urine, and blood/serum, contain multiple components that can contribute to an increased background or reduced analyte signal. In this mini-review, we discuss progress on electrochemical sensing in complex biofluids. We first introduce the challenge of drug titration in the management of various health conditions and provide an overview of the motivation for improved therapeutic drug monitoring, including current limitations. We then review progress on pharmaceutical drug detection from these biofluids with a focus on sample preprocessing, electrode modification for signal amplification, and/or electrode passivation to minimize fouling. Finally, we highlight promising strategies that have enabled robust drug quantification for clinical relevance and that may be useful for field-use systems.

1. Introduction

The electrochemical sensing of pharmaceuticals is an active area of research and development with multiple applications including environmental monitoring, pharmaceutical quality control, and diagnostics and therapeutic monitoring. Numerous previous reviews have described advances towards field-use systems for analyte monitoring. Sanavio and Krol (2015) described advances in the development of various point-of-care platforms as well as nanomaterials that could be used for therapeutic drug monitoring applications [1]. More recently, Sardini et al., (2020) reviewed printed electrochemical biosensors, i.e., sensors that use biological recognition elements, for a variety of target analytes including proteins, cells, and metabolites [2]. Zabihollahpoor et al., (2020) reviewed electrochemical sensors targeting antiseizure drugs for therapeutic drug monitoring [3]. In a complementary effort, Pollard et al., (2021) focused on electrochemical biosensors for therapeutic drug monitoring with an emphasis on promising sensing strategies that support continuous monitoring [4]. In a similar vein, Mobed et al., (2022) reviewed biosensors that specifically target antiseizure drugs [5]. Ozbek et al., (2022) reviewed potentiometric electrochemical sensors demonstrated using biological fluids and emphasized the limit of detection, recovery, and response time [6]. Kaur et al., (2022) [7] focused on electrochemical sensors using carbon-based nanomaterials for sensing drug and food pollutants. Most recently, Smith et al., (2023) [8] reviewed advances in the broad field of wearable sensing, while Calevilla et al., (2023) [9] reviewed work on sensors and biosensors applied to the monitoring of drugs for treating depression.
Here, we summarize recent progress (in the last five years) on the electrochemical sensing of pharmaceutical drugs in complex biofluids with an emphasis on addressing the challenge of detection in minimally processed samples. We first describe the motivation for the need for pharmaceutical monitoring by describing the challenge of drug titration in the management of many health conditions. Next, we provide an overview of the current practice of therapeutic drug monitoring, including specific drug classes of emphasis and describe the need for improved drug monitoring that would be enabled by field-use tools that can measure pharmaceuticals or their metabolic byproducts in complex biofluids. We then review progress in pharmaceutical drug detection from complex biofluids. We focus on sample preprocessing, electrode modification for signal amplification, and/or electrode passivation to minimize fouling and highlight the most promising strategies that would be useful for enabling personalized therapeutic drug monitoring in field-use systems. Finally, we close with a discussion of ongoing challenges in the achievement of point-of-care pharmaceutical monitoring systems.

2. Background

A major challenge in the management of health conditions that are treated using drug therapy is drug titration to control symptoms while minimizing adverse effects [10,11]. Drug titration can be further complicated in drugs with variable pharmacokinetics (PK) and strong interactions with other drugs. As an example, in the case of epilepsy, a neurological disease characterized by seizures that affects 65 million people worldwide [12], many of the antiseizure medications have serious toxicity risks, variable PK, and significant drug–drug interactions [13,14]. Thus, dose optimization can be difficult to achieve and may result in a reduced patient quality of life [15,16,17]. Drug overtreatment (dosing with higher than optimal drug levels in order to ensure seizure suppression) can further degrade patients’ quality of life, as well as lead to chronic health effects due to long-term drug use and/or medication noncompliance [18].
Therapeutic drug monitoring (TDM) has been discussed (and occasionally practiced) since the 1960s for multiple health conditions [10], including epilepsy [19]. It is based on the premise that drugs have a well-defined therapeutic range that can be broadly applied across the population. Although some clinicians currently use lab-based, blood-based TDM (Figure 1, outer cycle), the dependence on clinic sampling and lab analysis has limited its use to infrequent sampling. Further, the practice of TDM, which requires invasive venipuncture performed during a clinic visit and subsequent analyte quantification in a centralized lab, is a multi-day process that does not enable correlation of the patient drug level with their dosing schedule or experience with disease symptoms and adverse drug side effects. Prior work also indicates that substantial sex-based differences in PK can lead to adverse effects in women dosed using the established therapeutic ranges that were based on clinical trials composed mainly of men [20,21]. Likely as a result, the rate of adverse drug events has been reported to be much higher for women than men [20]. Finally, for each health condition, there are often specific subsets of patients who would benefit the most from multi-time point informed TDM. For example, women with epilepsy who are of childbearing age and, in particular, pregnant epilepsy patients, are high-risk populations for adverse drugs effects for themselves and the fetus [22,23,24].
In addition to therapeutic drug monitoring of antiseizure medication for epilepsy, there are numerous other classes of pharmaceuticals for which having a companion diagnostic device for drug monitoring could be beneficial. These include drugs used to treat other neurological conditions including Parkinson’s disease [25], drugs used in psychiatry [25] including those used to treat depression [26], drugs for the treatment of bacterial infections [27], and drugs for the treatment of viral infections such as hepatitis C [28]. Table 1 contains a summary of the classes of pharmaceuticals that are highlighted in this mini-review.
The use of noninvasive sampling fluids and point-of-care analysis could revolutionize TDM [48] and enable personalized TDM for each individual (Figure 1, inner cycle). Saliva [49,50], sweat [51], interstitial fluid [48], and urine [52], as alternatives to the gold-standard fluids of serum and blood, hold exceptional promise for point-of-care monitoring applications. For example, in the context of epilepsy management, the use of saliva as an alternative fluid for drug monitoring has been extensively discussed [53,54]. Critically, for monitoring applications in alternative biofluids, the drug concentration in the serum should be well-correlated with its concentration in the alternative biofluid. For example, in the context of antiseizure medication for epilepsy patients, the saliva concentration correlates strongly with the serum-free concentration at pharmacologically relevant levels [53,55,56,57,58]. Given the straightforward, noninvasive nature of saliva, sweat, interstitial fluid, and urine collection, monitoring using these biofluids could enable multi-point, patient-controlled sampling to more comprehensively inform the patient’s individual therapeutic drug levels [1,53,54] and potentially improve healthcare.
We have chosen to organize this mini-review in sections based on the type of biofluid, since each biofluid has unique advantages and disadvantages related to its composition, variability across the population, volume, and accessibility. Table 2 contains a summary of the main components of each biofluid and their advantages and disadvantages.

3. Advances in the Electrochemical Detection of Pharmaceuticals in Blood-Based Biofluids

Advances have been made in electrochemical sensing in blood and serum/plasma, and these studies are summarized in Table 3. From the standpoint of pharmaceutical monitoring, serum has the advantage of being the gold-standard matrix for drug detection, and serum is considered to be relatively uniform across individuals. The disadvantages of blood-based biofluids are that the collection of moderate volumes requires venipuncture in a clinic. Further, processing serum/plasma from whole blood generally requires lab-based processing, including centrifugation. As an alternative to small-volume-requirement sensors, fingerstick blood volumes (20 µL) can be compatible with collection at the point of care, and field-use tools (e.g., commercially available membranes) exist for processing small volumes of whole blood to plasma.
Most of the studies reviewed on electrochemical sensing in blood-based biofluids relied on high levels of dilution (greater than five-fold and sometimes upwards of 100-fold) or separation of the serum protein content before the electrochemical analysis. Only one of the studies, which focused on a field-use sensor, demonstrated detection from a finger-stick blood sample that would be compatible for sampling at the point of care [61]. However, the sensor still required dilution of the whole blood sample by 20-fold before analysis [61]. Thus, the ability to detect the target drug analyte directly in a complex matrix was not emphasized in the set of blood-based studies reviewed.
Progress in the electrochemical sensing of drug targets in blood-based biofluids has been made on other fronts, including signal enhancements using novel and more traditional electrode modifications, as well as the simultaneous monitoring of multiple target analytes (e.g., multiple drugs and/or potential interferents [62]). Given our focus on electrochemical monitoring to support drug titration, for each of the studies reviewed in Table 3, we highlight the drug target(s), the electrochemical method with the base electrode and sensing mechanism, the complex biofluid and any associated processing performed, the main cited strategies to improve the electrochemical signal, and the detection performance (e.g., limit of detection, dynamic range, and recoveries). Most of the sensors reviewed in this section used conventional carbon-based working electrodes composed of glassy carbon or carbon paste and then subsequently modified with nanostructures to enhance the electrochemical signal. Strategies for signal enhancement included the use of carbon-based materials such as multi-walled carbon nanotubes (well-known for having favorable properties such as a large surface area, high electrical conductivity, and resistance to fouling [63]) and graphene (also known for its large surface area, electrical conductivity, and mechanical strength [63]) in combination with other nanostructures. The nanostructures used varied from single metal nanoparticles composed of gold or cobalt to bi-metallic combinations of platinum-nickel and platinum with palladium to metal oxide nanoparticles such as zinc oxide and zirconium oxide. Further, more complex, novel combinations of layered structures were used for signal enhancement that included ionic liquid crystals and molecularly imprinted polymers. These modifications were utilized to generate robust electrochemical signals acquired using a variety of electrochemical methods, but most often via differential pulse voltammetry (DPV) or square-wave voltammetry (SWV) for drug target quantification.
Examples of notable signal enhancement strategies used in blood-based diagnostics are highlighted in Figure 2. Ibrahim et al. [64] demonstrated the quantitative detection of cyproterone acetate, an anti-androgen drug used in cancer treatment. Their sensor used a base electrode composed of glassy carbon paste that was then modified with a composite of multi-walled carbon nanotubes and gold nanoparticles. They optimized the concentrations of gold nanoparticles and multi-walled carbon nanotubes to achieve a substantial increase in the signal over the untreated electrode case (Figure 2A). Shalauddin et al. [65] targeted diclofenac sodium, an analgesic and anti-inflammatory drug used to treat arthritis and other conditions. In their sensor, they modified a base glassy carbon electrode with a composite of nanocellulose and multi-walled carbon nanotubes. They demonstrated significant gains in the signal using a combination of nanocellulose and multi-walled carbon nanotubes compared to using each component alone (Figure 2B). Atta et al. [62] demonstrated the simultaneous detection of multiple small molecule analytes, dobutamine, and amlodipine for the treatment of cardiac conditions, as well as acetaminophen and ascorbic acid. They modified a base glassy carbon electrode with a composite of multi-walled carbon nanotubes, ionic liquid crystals, graphene, and 18-Crown-6. Using differential pulse voltammetry, they showed well-resolved detection of the four different small molecule targets (Figure 2C).
Table 3. Summary of advances in drug detection in blood and serum. SWV, square-wave voltammetry; DPV, differential pulse voltammetry; CA, chronoamperometry; CV, cyclic voltammetry; AdSSWV, adsorptive stripping square-wave voltammetry; LOD, limit of detection; DR, dynamic range.
Table 3. Summary of advances in drug detection in blood and serum. SWV, square-wave voltammetry; DPV, differential pulse voltammetry; CA, chronoamperometry; CV, cyclic voltammetry; AdSSWV, adsorptive stripping square-wave voltammetry; LOD, limit of detection; DR, dynamic range.
Drug and Health ConditionElectrochemical Method/Base Working Electrode/Sensing MechanismComplex BiofluidStrategies to Improve Electrochemical SignalPerformance Metrics (Note that Metrics are Provided for Each of the Targets in the Order Listed in Column 1)Ref.
Carbamazepine to treat epilepsy seizuresSWV; gold electrode;
aptamer binding to carbamazepine reduces the distance between the methylene blue tag and the electrode and increases the current signal
Fingerstick blood-diluted 20-foldHigh packing density of aptamer for increased sensitivity to the targetLOD 2.1 nM in serum for 5 min and linearly over the 17 to 51 µM therapeutic range[61]
Daclatasvir, sofosbuvir, and ledipasvir for the treatment of hepatitis CDPV; glassy carbon electrodes; oxidation of each of the drugs Serum-diluted 200-foldMulti-walled carbon nanotubes in an ionic liquid crystal and cobalt nanoparticlesLODs 1.9 nM, 7.3 nM, 0.28 nM; linear DRs of 0.07 to 1 µM and 5 to 50 µM, 0.3 to 8 µM and 10 to 100 µM, 0.02 to 1 µM and 3 to 100 µM; recoveries of 99.7 to 102.8% for ledipasvir[66]
Cyproterone acetate for the treatment of prostate cancerSWV; glassy carbon paste electrodes; reduction of cyproterone acetateSerum-proteins separated out with ethanol precipitation and centrifugationMulti-walled carbon nanotubes and gold nanoparticlesLOD 17.7 nM; linear DR of 99 nM to 8.3 µM; sensitivity 117 µA/µM per cm2[64]
Regorafenib for the treatment of hepatocellular carcinomaDPV; glassy carbon electrodes; oxidation of regorafenib Serum-diluted 50-foldZirconium oxide nanoparticles and reduced graphene oxideLOD 17 nM in buffer (not reported in serum); linear DR of 11 to 343 nM in buffer; recoveries of 97.2 to 102.6%[40]
Doxorubicin and dasatinib for treatment of breast cancerCA and SWV; carbon paste electrodes; oxidation of drugsSerum-handling was not described Zinc oxide nanoparticles and butyl-3-methylimidazolium tetrafluoroborate and liquid paraffinLOD 9 nM and 0.5 µM in buffer (not reported in serum); linear DR of 0.07 to 500 µM, 9.0 nM to 0.5 µM in buffer; recoveries of 98.1 to 102.3%[67]
Acetaminophen and etilefrine (with dopamine)DPV; glassy carbon electrodes; oxidation of each of small moleculeSerum-10-fold dilution followed by another 125-fold dilutionPlatinum-nickel nanoparticles and reduced graphene oxideLOD 8.2 µM, 14.9 µM, 0.0025 µM in buffer (not reported in serum); linear DR of 4.0 to 60 µM, 4.0 to 100 µM, 0.05 to 0.5 µM in buffer; recoveries of 95 to 108%[68]
Dobutamine and amlodipine for the treatment of cardiac issues and acetaminophen and ascorbic acidDPV; glassy carbon electrodes; oxidation of each speciesSerum-20-fold dilution Composite of multi-walled carbon nanotubes, ionic liquid crystal, graphene and 18-Crown-6 enables the simultaneous detection of four speciesLOD 0.50 nM, 0.14 nM, 0.09 nM, 9.2 nM; linear DR of 0.02 to 40 µM, 0.008 to 30 µM, 0.001 to 20 µM, 0.4 to 40 µM; recoveries of 97.1 to 102.7%[62]
Acetaminophen (with tryptophan and caffeine)DPV; glassy carbon electrodes; oxidation of each speciesSerum-25-fold dilutionTin sulfide and titanium dioxide on graphene oxide sheetsLOD 7.5 nM, 7.8 nM, 4.4 nM in buffer (not reported in serum); linear DR of 9.8 nM to 280 µM, 13 nM to 157 µM, 16 nM to 333 µM in buffer; recoveries of 98% and 99% for acetaminophen[69]
Diclofenac sodium as an analgesic and anti-inflammatory for arthritis and other conditionsDPV and CV; glassy carbon electrodes; oxidation of diclofenac sodiumSerum-filtered and diluted 10-foldNanocellulose and multi-walled carbon nanotubesLOD 0.12 µM in buffer (not reported in serum); linear DR of 0.05 to 1 µM in buffer; recoveries of 99.3 to 102.0% [65]
Doxorubicin and dasatinib for the treatment of breast cancerAdSSWV; glassy carbon electrodes; oxidation of each drugSerum-10-fold dilutionPalladium and platinum nanoparticles with multi-walled carbon nanotubesLOD 0.86 nM, 6.72 nM in buffer (not reported in serum); linear DR of 4.4 nM to 8.6 µM, 38 nM to 9.9 µM in buffer; recoveries of 99.1 to 100.6%[70]
N-acetylcysteine for multiple indicationsDPV; carbon paste electrodes; oxidation of each drugSerum-10-fold or greaterSilica nanoparticles and boron trifluoride and 4,4′-dihydroxybiphenylLOD 0.33 µM in buffer (not reported in serum); linear DRs of 1.0 to 41.5 µM and 41.5 to 101.5 µM in buffer; agreed to within 1% of HPLC[71]
Chloroquine to treat malaria, rheumatoid arthritis, and cancerCV and DPV; glassy carbon electrodes; oxidation of chloroquineSerum-5-fold dilutionTungsten disulfide quantum dots with reduced graphene oxideLOD 0.04 µM; linear DR of 0.5 to 82 µM[72]
Olanzapine for the treatment of schizophreniaPotentiometric measurement and carbon paste electrodesSerum-10-fold dilutionOlanzapine-
tungstophosphate
LOD 0.5 µM in buffer (not reported in serum); linear DR of 0.75 to 560 µM in buffer; recoveries of 97.8 to 101.6%[39]
Epinephrine to treat allergic reactions, cardiac arrest, and hypertensionDPV; glassy carbon electrodes; oxidation of drugsSerum-proteins separated out with ethanol precipitation and centrifugationZinc oxide nanoparticles and multi-walled carbon nanotubesLOD 0.016 µM in buffer (not reported in serum); linear DR of 0.4 to 2.4 µM in buffer; recoveries of 100.4 to 101.3%[41]
Azithromycin for the treatment of bacterial infectionsDPV; glassy carbon electrodes; oxidation of azithromycinPlasma-filtered using 0.45 µm filter and diluted 10-foldMolecularly imprinted polymer LOD 0.85 nM in buffer (not reported in serum); linear DR of 13 nM to 67 µM in buffer; recovery of 102.4%[73]
Epirubicin and methotrexate for breast cancer treatmentDPV; glassy carbon electrodes; oxidation of each drugSerum-filtered using 0.45 µm filter and diluted 5-foldZinc oxide nanoflowers doped with cerium LOD 2.3 nM, 6.3 nM in buffer (not reported in serum); linear DR of 0.01 to 600 µM, 0.01 to 500 µM in buffer; recoveries of 98.0 to 102.3%[74]
Rifampicin to treat bacterial infectionsDPV; glassy carbon electrodes; oxidation of drugsSerum-indicated dilution of 3-fold Titanium dioxide nanoparticles on reduced graphene oxideLOD 0.03 µM in buffer (not reported in serum); linear DR of 0.01 to 0.1 nM in buffer; recoveries of 95 to 100%[75]
Levofloxacin for treating bacterial infectionsPotentiometric measurement; carbon paste electrodesSerum-diluted 25-foldPVC coatingLOD 10 µM in buffer (not reported in serum); linear DR of 10−2 to 10−4 M in buffer; recoveries of 95.6 to 98.7% for CPE[76]
Mefenamic acid, a non-steroidal anti-inflammatory drugCV and DPV; carbon paste electrodes; oxidation of mefenamic acidSerum-handling not describedCopper vanadium oxide nanostructures
(Cu5V2O10)
LOD 2.3 nM in buffer (not reported in serum); linear DR of 0.01 to 470 µM in buffer; recoveries of 98.3 to 110%[77]
Mefenamic acid, a non-steroidal anti-inflammatory drugCV and DPV; carbon paste electrodes; oxidation of mefenamic acidSerum-handling not describedTerbium titanate nanostructures (Tb2Ti2O7)LOD 2.4 nM in buffer (not reported in serum); linear DR of 0.01 to 400 µM in buffer; recoveries of 92.0 to 107%[78]
Complementary to the work performed using carbon-based electrodes, Chung et al. [61] described the development of an aptamer-based sensor for carbamazepine, an antiseizure drug that is widely prescribed for epilepsy. Their sensor used gold electrodes onto which an aptamer specific to carbamazepine was attached via a thiol [61]. The aptamer was tagged with methylene blue, such that upon the binding of the aptamer to the drug target, the methylene blue was localized close to the electrode and increased the electron transfer rate and current signal [61]. A key feature of their sensor was the high packing density of the aptamer to support sensitive drug detection. The authors demonstrated the assay selectivity against interference by analogues of carbamazepine and carbamazepine quantification within 5 min in diluted fingerstick blood (see Figure 3A).

4. Advances in the Electrochemical Detection of Pharmaceuticals in Alternative Fluids

Advances have also been made in electrochemical sensing in alternative biofluids to blood/serum/plasma, and these studies are summarized in Table 4. In the following discussion, we further highlight the strategies described that have supported robust drug quantification in alternative biofluids with an emphasis on methods that address unique challenges presented by complex biofluids, including an enhanced electrochemical signal against a high background, distinguishing the target signal in the presence of specific interference, and the prevention of electrode fouling.

4.1. Detection of Analyte Drugs in Saliva

Saliva has long been considered a promising biofluid for analyte detection due to its ease of collection, i.e., a pain-free, rapid process that can be performed ‘on demand’. However, saliva poses multiple challenges that have hindered its broader utilization in TDM. Saliva has a complex composition that varies across individuals in the population, as well as within an individual from day-to-day. Saliva is mainly composed of water and ions that give it a buffering capacity [59]; ion concentrations can change based on the saliva flow rate, e.g., whether the saliva is stimulated or unstimulated [59]. Saliva contains many different proteins, and their concentrations may vary based on the flow rate and physiological state of the cardiovascular system [59]. Saliva also contains hormones as well as small concentrations of uric acid, glucose, amino acids, lipids, and fatty acids [59]. Due to this complex composition of saliva and the associated potential for signal interference, the quantification of analytes in saliva can be challenging. Further, the analyte concentration in saliva is often at substantially lower levels than in the serum. Thus, strategies to enhance the target analyte signal over that of the background and interferents are critical for robust analyte quantification in saliva.
Several groups have reported different strategies to address the high background caused by interferents in saliva. Wentland et al. investigated a series of signal enhancement strategies to enable the detection of the antiseizure medication carbamazepine in commercially purchased pooled saliva [79]. They demonstrated that the use of the anionic surfactant sodium dodecyl sulfate (SDS) (as a facilitator of the interaction of carbamazepine with the electrode) combined with incubation resulted in a substantially increased carbamazepine signal compared to that with the bare stencil-printed carbon electrode [79]. In a follow up study, they demonstrated progress towards a field-use system composed of a disposable electrochemical flow cell (with all reagents incorporated dry) and a miniature potentiostat that enabled carbamazepine quantification at therapeutically relevant levels in saliva (see Figure 3B) [80]. Additional work is still needed to address the inter- and intra-individual variability of saliva.
In a complementary effort, Lin et al. [82] demonstrated the use of targeted surface engineering to create a ‘nondistorted potential window’ for the detection of the analyte acetaminophen in saliva with the interferents tyrosine, tryptophan, and uric acid. They accomplished this through a combination of hydrogen passivation of the base boron-doped diamond electrode and a nonfouling Nafion layer that both enhanced and shifted the acetaminophen signal [82]. The investigators noted the use of saliva centrifugation before analysis (i.e., the removal of cellular debris and higher-molecular-weight proteins from the saliva matrix), so the sensor robustness against unprocessed saliva was not determined [82]. The sensor was also demonstrated against a background of unprocessed sweat (see Figure 4A) [82].
Most recently, Gomes and Raymundo-Pereira demonstrated the detection of acetaminophen against a background of undiluted saliva using differential pulse voltammetry and screen-printed carbon electrodes (see Figure 3C) [81]. Notably, the authors showed the detection of their drug target against a background of undiluted saliva collected via the passive drool method [81]. Further they showed time course data of the drug level in saliva before and after taking a drug dose [81]. The authors attributed the high sensitivity of their sensor and its robustness to fouling to their electrode pretreatment with sulfuric acid [81] but did not quantify the difference in performance. A discussion of possible strategies for the implementation of the pretreatment step in a point-of-care device would have been useful.

4.2. Detection of Analyte Drugs in Sweat

Sweat is another biofluid for which there has been much interest due to its potential for noninvasive continuous sampling. Sweat is made up of 99% water, and the main components of the other 1% are electrolytes, fatty acids, lactic acid, and multiple nitrogen metabolites, including uric acid, urea, and ammonia [51]. Sweat also contains small amounts of proteins (including antibodies), peptides, amino acids, amines, metal ions, and ethanol [51]. De Castro et al. [51] have argued that sweat has high impact potential as a biofluid for use in drug sensing. The challenges of using sweat as a monitoring fluid include its low resting secretion rates (10 to 100 nL/min per cm2) and the need to minimize evaporation effects in order to obtain an adequate biofluid volume for accurate analysis [34].
Several groups have reported advances in the electrochemical detection of pharmaceuticals in sweat. Levodopa, a drug used in the treatment of Parkinson’s disease, has been a common target due to its short half-life and the fact that optimal dosing becomes more difficult as the disease progresses. The Javey group described the development of an on-the-arm wearable device for monitoring levodopa using amperometry (see Figure 4B) [83]. The device electrodes were composed of gold nanodendrites deposited on a thin gold/chromium layer and then modified by deposition of thionin acetate salts [83]. Further modifications to the working electrode included a layer of glutaraldehyde-crosslinked tyrosinase and then Nafion [83]. Levodopa quantification was achieved via enzyme-facilitated levodopa oxidation. The authors noted the key features of their sensor as being the use of gold nanodentrites to increase the surface area for enzyme attachment and the use of Nafion film as a nonfouling layer that provides sensor signal stability in the longer-term [83]. They demonstrated levodopa quantification of their sensor in the range of 0 to 20 µM against a background of sweat, as well as the monitoring capability of their sensor on human subjects who had ingested levodopa-containing fava beans and using simulated sweat via iontophoretic stimulation or exercise [83].
In a follow-up study, the Javey group further advanced sweat sensing by demonstrating a system to monitor levodopa in subjects at rest (see Figure 4C) [84]. A key feature of their sensor system that was critical in facilitating the collection of the very small volumes of sweat produced during the resting state was the human interface component composed of an agarose-glycerol hydrogel film on a PVA-coated SU-8 [84]. The sensor also included a Nafion coating to resist fouling [84]. Further, in order to provide accurate quantification of levodopa levels over time, their system included impedance-based sweat secretion rate monitoring using a configuration of interdigitated electrodes [84]. And, the miniature footprint of the system was also key to the achievement of conformal placement on the target areas of the fingertip and wrist while allowing for the performance of daily activities [84]. Against a background of sweat, the levodopa sensor was characterized as having improved sensitivity over their prior work, despite a decrease in the sensing area [84]. Finally, the investigators demonstrated on-the-body sweat monitoring that showed increasing levodopa levels with increasing doses of broad bean consumption [84].
The Wang group also reported the development of their touch-based levodopa sensor for sweat sensing [85]. Their sensor uses screen-printed carbon electrodes modified with crosslinked tyrosinase and amperometry [85]. Similar to the studies described above, the electrochemical sensor utilizes the tyrosinase-facilitated oxidation of levodopa. However, in this sensor design, the levodopa concentration is tracked via dopaquinone reduction after its conversion from levodopa [85]. The investigators noted that an advantage to their method is its robustness to fouling of the electrode via unintended quinone polymerization reactions [85]. Notably, simultaneous measurements of levodopa levels in capillary blood showed a correlation with the levodopa levels in nonstimulated sweat [85].
More recently, Raymundo-Pereira et al. [86] reported the development of a wearable sensor in glove format for monitoring the pharmaceuticals acetaminophen and paroxetine. The sensor was characterized for its mechanical stability, selectivity, and response in artificial sweat, as well as the reproducibility of a single drug concentration against a background of human sweat combined with artificial sweat [86]. The investigators noted the requirement of electrode pretreatment with an acidic solution in order to reduce impurities and achieve the desired sensor response [86].

4.3. Detection of Analyte Drugs in Interstitial Fluid

Interstitial fluid (ISF) is another complex matrix for which there is much interest due to its potential for noninvasive continuous sampling. The main components of ISF are amino acids, carbohydrates, and fatty acids [48]. Further, drugs and small micronutrients are able to pass through the capillary walls into the ISF [48]. Aside from the fact that it can be noninvasively sampled, ISF is preferred over blood samples because it does not contain proteins and other cells that can interfere with drug-level analysis [48], and thus, it does not require the extensive sample manipulations required to process whole blood to serum.
The two main challenges associated with sensing in ISF are the very small fluid volumes that must be accessed from the extracellular spaces and then directed to a micro-electrode system for detection and the resistance to biofouling of the electrodes over the timescale of monitoring with the microelectrodes. The Wang group reported on their promising microneedle-based electrochemical sensing platform in which carbon paste electrodes and a silver wire reference electrode are embedded into hollow microneedles [87]. Goud et al., applied their platform to the detection of levodopa by combining tyrosinase with the carbon paste in order to monitor the enzyme-facilitated oxidation of levodopa using amperometry [88]. In addition, a Nafion coating was included to provide resistance to interference by negative species [88] and to reduce electrode fouling by larger-molecular-weight proteins [63].
In a follow up study, the Wang group demonstrated the detection of apomorphine, another drug used in the treatment of Parkinson’s disease [89]. The investigators demonstrated that doping the carbon paste working electrode with rhodium nanoparticles resulted in larger differences in the current peaks compared to the situation without doping and thus increased the resolution of the apomorphine concentrations of interest using artificial ISF within a gel as a test bed [89]. A nonfouling layer of Nafion was incorporated on the working electrode, and with it, the authors reported the successful detection of apomorphine over potential interferents including histidine, phenylamine, tryptophan, and acetaminophen [89] (see Figure 5A).
In a complementary effort, the Emaminejad group [90] demonstrated a microneedle-based electrochemical sensor for two antibiotics, tobramycin and vancomycin, with narrow therapeutic ranges and for which therapeutic drug monitoring has been indicated. Their wearable sensor uses gold-nanoparticle-coated acupuncture needles fixed within a PDMS substrate [90]. Aptamers, functionalized with the redox reporter methylene blue, were attached to the electrode surface, such that their conformational change during target binding resulted in a detectable change in the voltammetric signal [90]. The authors noted that their system was robust to biofouling, as demonstrated by the limited drift when exposed to a protein-spiked buffer over 14 h [90]. The highlights of the sensor’s capabilities include in vivo (rat) correlation studies of the drug concentration in serum vs. ISF [90] (see Figure 5B).

4.4. Detection of Analyte Drugs in Urine

Urine is a complex biological fluid, comprising water, urea, uric acid, inorganic salts, enzymes, nucleic acids, vitamins, proteins, amino acids, hormones, mesothelin, beta-microglobuin, urokinase, antibiotics, and mycomycin [52]. Urine collection has the advantage of being noninvasive and easy to sample, and there may be a longer time window available for drug detection compared to other biological matrices such as saliva or blood [60]. However, urine samples cannot be collected ‘on demand’ and may be contaminated if the collection is not conducted carefully [60].
As was the case for the studies reviewed on blood-based biofluids, the majority of studies reporting on electrochemical drug sensing in urine have relied heavily on processing to remove the protein content or sample dilution in order to mitigate interference effects. Only one study by Ishii et al. demonstrated drug quantification in unprocessed urine [91]. In that work, the investigators used boron-doped diamond (BDD) electrodes to monitor the reduction of the diuretic triamterene. The rationale for the choice of the BDD electrode include its resistance to biofouling, its stability, and its relatively large potential window [91]. The choice to monitor the triamterene reduction was motivated by the occurrence of high background current at higher potentials. When assessing the drug content in individual urine samples, high levels of ascorbic acid were found to be problematic, as they interfered with the triamterene signal [91]. Also noteworthy was the variability of the triamterene potential in the different individual urine samples and the potential need for more sophisticated signal analysis algorithms.
Notable signal enhancement strategies used in studies on urine-based electrochemical sensing overlap with the studies highlighted using blood-based biofluids and include the use of multi-walled carbon nanotubes combined with gold nanoparticles [64], palladium and platinum nanoparticles [70], or nanocellulose [65]. Other reported electrode modifications for enhanced electrochemical signals include platinum nanoflowers with reduced graphene oxide [92] and a composite of iron oxide and polypyrrole and palladium [93]. In addition, multiple reports describe the use of molecularly imprinted polymers [73,94,95,96].
Table 4. Summary of advances in drug detection in alternative complex nonblood biological matrices. SWV, square-wave voltammetry; DPV, differential pulse voltammetry; CA, chronoamperometry; CV, cyclic voltammetry; AdSSWV, adsorptive stripping square-wave voltammetry; LOD, limit of detection; QR, quantitative resolution; DR, dynamic range.
Table 4. Summary of advances in drug detection in alternative complex nonblood biological matrices. SWV, square-wave voltammetry; DPV, differential pulse voltammetry; CA, chronoamperometry; CV, cyclic voltammetry; AdSSWV, adsorptive stripping square-wave voltammetry; LOD, limit of detection; QR, quantitative resolution; DR, dynamic range.
Drug and Health ConditionElectrochemical Method/Base Working Electrode/Sensing MechanismComplex BiofluidStrategies to Improve Electrochemical SignalPerformance Metrics (Note that Metrics are Provided for Each of the Targets in the Order Listed in Column 1)Ref.
Interferon gamma for treating cancer and infectionsAmperometry; screen-printed carbon electrodes treated with p-ABA diazonium salt to immobilize capture Ab; reduction of benzoquinone from an enzymatic reaction of label HRP, hydroquinone, and H2O2Saliva collected with a Salivette and then extracted using centrifugationOptimization of parameters including the capture Ab concentration and the concentrations of detected Ab and enzymes LOD 1.6 pg/mL in buffer; linear DR of 2.5 to 2000 pg/mL in buffer; in saliva, measurements agreed with ELISA to within 3%[97]
Carbamazepine to treat epilepsy seizuresSWV;
stencil-printed carbon electrodes; carbamazepine oxidation
Saliva—pooled, commercially purchased Sodium dodecyl sulfate in solution LOD 1 µM; average QR of 0.85 µM from 0 to 15 µM[79]
SWV;
stencil-printed carbon electrodes; carbamazepine oxidation
Saliva—pooled, commercially purchasedSodium dodecyl sulfate film on electrodesLOD 1 µM; average QR of 1.6 µM from 0 to 15 µM for field-use format sensor[80]
Acetaminophen/
paracetamol as an analgesic
DPV; oxygen-terminated boron-doped diamond electrode; oxidation of acetaminophenSaliva and sweat—saliva was processed by centrifugation before useHydrogen-
terminated boron-doped diamond electrode with a Nafion layer
LOD 1 µM; strong correlation in saliva (R2 = 0.92) and sweat (R2 = 0.95) with LC-MS/MS[82]
Acetaminophen/
paracetamol to manage pain
DPV; screen-printed carbon electrodes; oxidation of acetaminophenSaliva, unprocessedElectrochemical pretreatment consisting of cyclic voltammetry of 0.5 M sulfuric acid increased the electrode conductivity and signalLOD 14.5 µM; linear DR 25 to 150 µM[81]
Benzodiazepine for the treatment of depression, anxiety, and insomnia DPV; laser-scribed graphene electrodes functionalized with Ab capture; oxidation current change with Ab-Ag bindingSaliva collected with a swab and extracted by centrifugationOptimization of capture of the Ab concentration and blocking agent treatmentLOD 9.7 ng/mL in buffer; DR of 1.0 pg/mL to 500 ng/mL in buffer; simultaneous detection with amphetamine and cocaine in saliva [98]
Levodopa for the treatment of Parkinson’s diseaseAmperometry; gold-coated electrodes with tyrosinase; tyrosinase-catalyzed oxidation of levodopaSimulated sweat using iontophoretic stimulation or exerciseGold nanodendrite structures on gold electrodes and Nafion filmLOD 1 µM; DR of 0 to 20 µM; sensitivity of 17 nA/µM[83]
Amperometry; gold-coated electrodes with tyrosinase; tyrosinase-catalyzed oxidation of levodopaSweat generated at restGold nanodendrite structures on the gold electrodes and the Nafion-TBAB filmLOD 3 µM in buffer; linear DR of 0 to 50 µM in buffer; in situ sweat analysis[84]
Levodopa for treatment of Parkinson’s diseaseAmperometry; screen-printed carbon electrodes coated with crosslinked tyrosinase; dopaquinone reductionFingertip sweat using a touch sensorNote that an advantage to their method is its robustness to fouling of the electrode via unintended quinone polymerization reactionsLOD 300 nM in buffer; linear DR of 1 to 30 µM in buffer; in situ sweat analysis[85]
Acetaminophen/
paracetamol to manage pain and paroxetine as an antidepressant
DPV; screen-printed carbon electrodesArtificial sweat and human sweat combined with artificial sweat in equal partsPretreatment consisting of cyclic voltammetry of 0.5 M sulfuric acidLOD 0.25 µM, 0.49 µM in artificial sweat; recoveries of 106% and 112% in artificial sweat[86]
Levodopa for treatment of Parkinson’s diseaseSWV and CA; carbon paste electrodes with tyrosinase in microneedles; oxidation of levodopaArtificial interstitial fluidNafionLOD 0.5 µM in artificial ISF; linear DR of 0.5 to 3 µM in artificial ISF[88]
Apomorphine for treatment of Parkinson’s diseaseSWV and CA;
carbon paste electrodes in microneedles;
apomorphine oxidation
Artificial interstitial fluid containing protein interferents 2% rhodium nanoparticles and 1% Nafion film for stability against protein interferentsLOD 0.6 µM/0.75 µM (SWA/CA) in buffer; linear DR of 10 to 60 µM in the skin mimic model; sensitivity of 3.8 nA/µM in the skin mimic model[89]
Tobramycin and vancomycin for bacterial infectionsSWV; gold coating of an acupuncture needle; the binding of aptamer to the antibiotic reduces the distance between the methylene blue tag and the electrode and increases the current signal Interstitial fluid— in vivo on ratGold nanoparticle coating enhances the signal compared to evaporated gold filmResponse curves from 0 to 100 µM for each in artificial ISF; correlation between blood and ISF tobramycin levels in mice[90]
Cyproterone acetate for the treatment of prostate cancerSWV; glassy carbon paste electrodes; reduction of cyproterone acetateUrine—diluted with bufferMulti-walled carbon nanotubes and gold nanoparticlesLOD 17.9 nM; linear DR of 99 nM to 5.0 µM; sensitivity 173 µA/µM per cm2[64]
Diclofenac sodium as an analgesic and anti-inflammatory for arthritis and other conditionsDPV; glassy carbon electrodes; oxidation of diclofenac sodium Urine—diluted 4-foldNanocellulose and multi-walled carbon nanotubesLOD 0.12 µM in buffer (not reported in urine); linear DR of 0.05 to 1 µM in buffer; recoveries of 98.0 to 104.0%[65]
Doxorubicin and dasatinib for the treatment of breast cancerAdSSWV; glassy carbon electrodes; oxidation of each drugUrine—10-fold dilutionPalladium and platinum nanoparticles with multi-walled carbon nanotubesLOD 0.86 nM, 6.72 nM in buffer (not reported in urine); linear DR of 4.4 nM to 8.6 µM, 38 nM to 9.9 µM in buffer; recoveries of 98.8 to 99.5%[70]
Diclofenac sodium as an analgesic and anti-inflammatory for arthritis and other conditionsDPV; screen-printed carbon electrodes; oxidation of diclofenac sodium Urine—centrifuged to remove solidsPlatinum nanoflowers with reduced graphene oxide facilitated additional analyte on the electrode and improved electron transferLOD 40 nM in buffer (not reported in urine); linear DR 0.1 to 100 µM in buffer; recoveries of 84 to 105%[92]
Azithromycin for the treatment of bacterial infectionsDPV; glassy carbon electrodes; oxidation of azithromycinUrine and tears— filtered using an 0.45 µm filter and diluted 10-foldMolecularly imprinted polymerLOD 0.85 nM in buffer (not reported in urine); linear DR of 13 nM to 67 µM in buffer; recoveries of 98.0 to 106.3%[73]
Epirubicin and methotrexate for breast cancer treatmentDPV; glassy carbon electrodes; oxidation of each drugUrine—filtered using 0.45 µm filter and diluted 5-foldZinc oxide nanoflowers doped with ceriumLOD 2.3 nM, 6.3 nM in buffer (not reported in urine); linear DR of 0.01 to 600 µM, 0.01 to 500 µM in buffer; recoveries of over 97.1 to 102.6%[74]
Levofloxacin for treating bacterial infectionsPotentiometric measurement; carbon paste electrodesUrine—diluted 25-foldPVC coatingLOD 10 µM in buffer (not reported in urine); linear DR of 10−2 to 10−4 M in buffer; recoveries of 94.5 to 98.4% for CPE[76]
Triamterene as a diureticCV, CA, SWV;
boron-doped diamond electrodes; reduction of triamterene
Pooled and individual urineNote that the electrode choice provides resistance to biofouling, stability, and a relatively large potential windowLOD 7.80 nM pooled and 20.8 nM individual urine[91]
Nalbuphine as an analgesicPotentiometric measurement; screen-printed carbon electrodesUrine—diluted 10-foldComposite of polyaniline with multi-walled carbon nanotubes and PVC with
molecularly imprinted polymer beads
LOD 11 µM in buffer (not reported in urine); linear DR of 43 to 3300 µM in buffer; recoveries of 91.0 to 101.5%[94]
Methotrexate for cancer treatment DPV; screen-printed graphite electrodes; oxidation of methotrexate and folic acidUrine—centrifuged, supernatant filtered using 0.45 µm filter and diluted at least 2.5-foldComposite of iron oxide and polypyrrole and palladiumLOD 7.0 nM in buffer (not reported in urine); linear DR of 0.03 to 100 µM in buffer; recoveries of 97.8 to 103.1%[93]
Sulfanilamide for the treatment of bacterial infectionsCV and DPV; 3D printed carbon black-PLA electrodes; oxidation of sulfanilamideArtificial urine—diluted 10-fold Electrode pretreatment of NaOH solution at 1.4 V and −1 V for 200 s eachLOD 12 nM in buffer; linear DR of 1 to 39 µM in buffer; recoveries of 99.1 to 101.9% in synthetic urine[99]
Trimethoprim for the treatment of bacterial infectionsDPV; carbon paste electrodes with iron oxide and multi-walled carbon nanotubes; oxidation of trimethoprimUrine—centrifuged and supernatant analyzedLayered structure consisting of base electrode material and reduced graphene oxide and
molecularly imprinted polymer with iron oxide and multi-walled carbon nanotubes
LOD 1.2 nM in buffer (not reported in urine); linear DRs of 0.004 to 0.08 µM and 0.08 to 500 µM in buffer; recoveries of 95.0 to 110.0%[95]
Aminophylline for the treatment of bronchial asthmaDPV; glassy carbon electrodes; oxidation of aminophyllineUrine—filtered and dilutedMolecularly imprinted polymer and graphene oxideLOD 2.1 pM in buffer (not reported in urine); linear DR of 37 pM to 1 mM in buffer; recoveries of 98.2 to 99.6%[96]
Ketoconazole for the treatment of fungal infectionsDPV, CV, CA; carbon paste electrodes; oxidation of ketoconazoleUrine—centrifuged, filtered, and dilutedMetal–organic framework composed of cerium and 1,3,5 benzene tricarboxylic acid and ionic liquidLOD 0.04 µM in buffer; linear DR of 0.1 to 110 µM in buffer; recoveries of 96.7 to 102.0%[100]

5. Summary and Ongoing Challenges

There has been exciting progress in the development of electrochemical sensing methods in a variety of biofluids. The studies reviewed investigated numerous drugs with strong clinical motivation for monitoring and have reported effective strategies for signal enhancement and mitigating the effects of interferents. In addition to electrode design to increase the conductivity and electroactive surface area, a key requirement for sensing in biofluids is the judicious choice of electrode materials/modifications to increase the fouling resistance and thus improve the signal stability. In the studies reviewed, choices for electrode modification often included multi-walled carbon nanotubes and graphene, materials that have been reported to have fouling resistance to particular species of interest [63]. Further, multiple studies have incorporated surfactant or Nafion coatings to mitigate electrode fouling by negatively charged species and/or larger-molecular weight-species such as proteins.
An unaddressed challenge in this area of electrochemical detection of pharmaceuticals is achieving robust analyte quantification from biofluids at the point of care. Most of the studies that performed measurements in blood and urine that were reviewed here relied on high levels of sample dilution (it is important to note that point-of-care detection was not necessarily a priority for all studies; some may have assumed the current status quo for patient sample collection in a clinic and subsequent lab-based analysis). Sample dilution could be implemented at the point of care, but would require either (i) additional user steps to manually perform the dilution or (ii) additional complexity (and cost) of the device in the form of an upstream sample processing module that automatically performs the dilution ([101] describes a microfluidic implementation demonstrated for blood processing and [102] describes H-filter enrichment of a drug over larger-molecular-weight interferents in the context of saliva processing). Further, sample dilution of the analyte would require a higher sensitivity point-of-care sensor, thus entailing the classic challenge of achieving high sensitivity in a field-use format.
For the studies that performed measurements in unprocessed biofluids, multiple challenges remain. Biofluid variability can be significant over time (e.g., daily for saliva) in an individual, as well as across the population. Thus, there is a need for more studies to assess electrochemical methods in raw samples without substantial processing as well as to validate methods using biofluid samples from many different individuals to assess the robustness more comprehensively.
Finally, the realization of effective companion diagnostics will require the thoughtful use of fabrication materials and methods that are compatible with field-use formats and that meet the requirements of usability, cost, maintenance, and performance. For example, with regard to fabrication, multiple articles cited in this review used either stencil- or screen-printed electrodes, methods which are consistent with a low cost per device and the ability to scale up. Additionally, progress has been reported in low-cost fabrication methods, including the use of stencil-printing materials other than carbon ink (e.g., glassy carbon [103]) and the use of the screen-printing process for modifications such as NafionTM [104]. These advances could be promising for incorporation in future work targeting electrochemical sensing for pharmaceutical detection at the point of care.

Author Contributions

Conceptualization, E.F. and K.K.; data curation, E.F., K.K. and N.L.; writing—original draft preparation, E.F. and N.L.; writing—review and editing, E.F., S.A.R., L.W., K.K., N.L. and M.L.J.; funding acquisition, E.F., M.L.J. and S.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the NIH/NIDCR under grant number R21DE031101 and the APC was funded by MDPI/Chemosensors. The content of this article is solely the responsibility of the authors and does not necessarily represent the official views of the NIH.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sanavio, B.; Krol, S. On the slow diffusion of point-of-care systems in therapeutic drug monitoring. Front. Bioeng. Biotechnol. 2015, 3, 30. [Google Scholar] [CrossRef] [PubMed]
  2. Sardini, E.; Serpelloni, M.; Tonello, S. Printed Electrochemical Biosensors: Opportunities and Metrological Challenges. Biosensors 2020, 10, 166. [Google Scholar] [CrossRef]
  3. Zabihollahpoor, A.; Rahimnejad, M.; Najafpour-Darzi, G.; Moghadamnia, A.A. Recent advances in electroanalytical methods for the therapeutic monitoring of antiepileptic drugs: A comprehensive review. J. Pharm. Biomed. Anal. 2020, 188, 113394. [Google Scholar] [CrossRef] [PubMed]
  4. Pollard, T.D.; Ong, J.J.; Goyanes, A.; Orlu, M.; Gaisford, S.; Elbadawi, M.; Basit, A.W. Electrochemical biosensors: A nexus for precision medicine. Drug Discov. Today 2021, 26, 69–79. [Google Scholar] [CrossRef]
  5. Mobed, A.; Shirafkan, M.; Charsouei, S.; Sadeghzadeh, J.; Ahmadalipour, A. Biosensors technology for anti-epileptic drugs. Clin. Chim. Acta 2022, 533, 175–182. [Google Scholar] [CrossRef]
  6. Ozbek, O.; Berkel, C.; Isildak, O. Applications of Potentiometric Sensors for the Determination of Drug Molecules in Biological Samples. Crit. Rev. Anal. Chem. 2022, 52, 768–779. [Google Scholar] [CrossRef] [PubMed]
  7. Kaur, H.; Siwal, S.S.; Chauhan, G.; Saini, A.K.; Kumari, A.; Thakur, V.K. Recent advances in electrochemical-based sensors amplified with carbon-based nanomaterials (CNMs) for sensing pharmaceutical and food pollutants. Chemosphere 2022, 304, 135182. [Google Scholar] [CrossRef]
  8. Smith, A.A.; Li, R.; Tse, Z.T.H. Reshaping healthcare with wearable biosensors. Sci. Rep. 2023, 13, 4998. [Google Scholar] [CrossRef]
  9. Caldevilla, R.; Morais, S.L.; Cruz, A.; Delerue-Matos, C.; Moreira, F.; Pacheco, J.G.; Santos, M.; Barroso, M.F. Electrochemical Chemically Based Sensors and Emerging Enzymatic Biosensors for Antidepressant Drug Detection: A Review. Int. J. Mol. Sci. 2023, 24, 8480. [Google Scholar] [CrossRef]
  10. Ensom, M.H.H.; Davis, G.A.; Cropp, C.D.; Ensom, R.J. Clinical pharmacokinetics in the 21st century—Does the evidence support definitive outcomes? Clin. Pharmacokinet. 1998, 34, 265–279. [Google Scholar] [CrossRef]
  11. Seiden, L.G.; Connor, G.S. The importance of drug titration in the management of patients with epilepsy. Epilepsy Behav. 2022, 128, 108517. [Google Scholar] [CrossRef] [PubMed]
  12. Kanner, A.M.; Bicchi, M.M. Antiseizure Medications for Adults with Epilepsy A Review. J. Am. Med. Assoc. 2022, 327, 1269–1281. [Google Scholar] [CrossRef]
  13. Perucca, E.; Meador, K.J. Adverse effects of antiepileptic drugs. Acta Neurol. Scand. 2005, 112, 30–35. [Google Scholar] [CrossRef]
  14. Perucca, P.; Gilliam, F.G. Adverse effects of antiepileptic drugs. Lancet Neurol. 2012, 11, 792–802. [Google Scholar] [CrossRef] [PubMed]
  15. St. Louis, E.K.; Minimizing, A.E.D. Adverse Effects: Improving Quality of Life in the Interictal State in Epilepsy Care. Curr. Neuropharmacol. 2009, 7, 106–114. [Google Scholar] [CrossRef] [PubMed]
  16. Mahrer-Imhof, R.; Jaggi, S.; Bonomo, A.; Hediger, H.; Eggenschwiler, P.; Kramer, G.; Oberholzer, E. Quality of life in adult patients with epilepsy and their family members. Seizure Eur. J. Epilepsy 2013, 22, 128–135. [Google Scholar] [CrossRef]
  17. Kubota, H.; Awaya, Y. Assessment of health-related quality of life and influencing factors using QOLIE-31 in Japanese patients with epilepsy. Epilepsy Behav. 2010, 18, 381–387. [Google Scholar] [CrossRef]
  18. Perucca, E. Overtreatment in epilepsy: Adverse consequences and mechanisms. Epilepsy Res. 2002, 52, 25–33. [Google Scholar] [CrossRef]
  19. Landmark, C.J.; Johannessen, S.I.; Patsalos, P.N. Therapeutic drug monitoring of antiepileptic drugs: Current status and future prospects. Expert Opin. Drug Metab. Toxicol. 2020, 16, 227–238. [Google Scholar] [CrossRef]
  20. Zucker, I.; Prendergast, B.J. Sex differences in pharmacokinetics predict adverse drug reactions in women. Biol. Sex Differ. 2020, 11, 32. [Google Scholar] [CrossRef]
  21. Soldin, O.P.; Mattison, D.R. Sex Differences in Pharmacokinetics and Pharmacodynamics. Clin. Pharmacokinet. 2009, 48, 143–157. [Google Scholar]
  22. Adab, N.; Chadwick, D.W. Management of women with epilepsy during pregnancy. Obstet. Gynaecol. 2006, 8, 20–25. [Google Scholar] [CrossRef]
  23. Patel, S.I.; Pennell, P.B. Management of epilepsy during pregnancy: An update. Ther. Adv. Neurol. Disord. 2016, 9, 118–129. [Google Scholar] [CrossRef]
  24. Thomas, S.V. Management of epilepsy and pregnancy. J. Postgrad. Med. 2006, 52, 57–64. [Google Scholar]
  25. Schoretsanitis, G.; Paulzen, M.; Unterecker, S.; Schwarz, M.; Conca, A.; Zernig, G.; Gründer, G.; Haen, E.; Baumann, P.; Bergemann, N.; et al. TDM in psychiatry and neurology: A comprehensive summary of the consensus guidelines for therapeutic drug monitoring in neuropsychopharmacology, update 2017, a tool for clinicians. World J. Biol. Psychiatry 2018, 19, 162–174. [Google Scholar] [CrossRef] [PubMed]
  26. Grundmann, M.; Kacirova, I.; Urinovska, R. Therapeutic monitoring of psychoactive drugs–Antidepressants: A review. Biomed. Pap. 2015, 159, 35–43. [Google Scholar] [CrossRef] [PubMed]
  27. Roberts, J.A.; Norris, R.; Paterson, D.L.; Martin, J.H. Therapeutic drug monitoring of antimicrobials. Br. J. Clin. Pharmacol. 2012, 73, 27–36. [Google Scholar] [PubMed]
  28. Mallon, P.W.G.; Ray, J.; Cooper, D.A. Effect of therapeutic drug monitoring on outcome in antiretroviral experienced HIV-infected individuals. J. Clin. Virol. 2003, 26, 223–227. [Google Scholar] [CrossRef]
  29. Pargas, C.D.; Elhessy, A.H.; Abouei, M.; Gesheff, M.G.; Conway, J.D. Tobramycin Blood Levels after Local Antibiotic Treatment of Bone and Soft Tissue Infection. Antibiotics 2022, 11, 336. [Google Scholar] [CrossRef]
  30. Boyle, A.; Ondo, W. Role of Apomorphine in the Treatment of Parkinson’s Disease. CNS Drugs 2015, 29, 83–89. [Google Scholar] [CrossRef]
  31. Oertel, W.H.; Quinn, N.P. Parkinson’s disease: Drug therapy. Bailliere’s Clin. Neurol. 1997, 6, 89–108. [Google Scholar]
  32. Almalki, R.S.; Eweis, H.; Kamal, F.; Kutbi, D. Methotrexate Toxicity: Molecular Mechanisms and Management. J. Pharm. Res. Int. 2021, 33, 204–217. [Google Scholar] [CrossRef]
  33. Pugazhendhi, A.; Edison, T.; Velmurugan, B.K.; Jacob, J.A.; Karuppusamy, I. Toxicity of Doxorubicin (Dox) to different experimental organ systems. Life Sci. 2018, 200, 26–30. [Google Scholar] [CrossRef] [PubMed]
  34. Brunner, A.M.; Costa, D.B.; Heist, R.S.; Garcia, E.; Lindeman, N.I.; Sholl, L.M.; Oxnard, G.R.; Johnson, B.E.; Hammerman, P.S. Treatment-Related Toxicities in a Phase II Trial of Dasatinib in Patients with Squamous Cell Carcinoma of the Lung. J. Thorac. Oncol. 2013, 8, 1434–1437. [Google Scholar] [CrossRef] [PubMed]
  35. Prado, C.M.M.; Lima, I.S.F.; Baracos, V.E.; Bies, R.R.; McCargar, L.J.; Reiman, T.; Mackey, J.R.; Kuzma, M.; Damaraju, V.L.; Sawyer, M.B. An exploratory study of body composition as a determinant of epirubicin pharmacokinetics and toxicity. Cancer Chemother. Pharmacol. 2011, 67, 93–101. [Google Scholar] [CrossRef] [PubMed]
  36. Giordano, N.; Nardi, P.; Santacroce, C.; Geraci, S.; Gennari, C. Acute hepatitis induced by cyproterone acetate. Ann. Pharmacother. 2001, 35, 1053–1055. [Google Scholar] [CrossRef]
  37. Sastre, J.; Argiles, G.; Benavides, M.; Feliu, J.; Garcia-Alfonso, P.; Garcia-Carbonero, R.; Grávalos, C.; Guillén-Ponce, C.; Martínez-Villacampa, M.; Pericay, C. Clinical management of regorafenib in the treatment of patients with advanced colorectal cancer. Clin. Transl. Oncol. 2014, 16, 942–953. [Google Scholar] [CrossRef]
  38. Burgess, S.; Partovi, N.; Yoshida, E.M.; Erb, S.R.; Azalgara, V.M.; Hussaini, T. Drug Interactions with Direct-Acting Antivirals for Hepatitis C: Implications for HIV and Transplant Patients. Ann. Pharmacother. 2015, 49, 674–687. [Google Scholar] [CrossRef]
  39. Rouhani, M.; Soleymanpour, A. A new selective carbon paste electrode for potentiometric analysis of olanzapine. Measurement 2019, 140, 472–478. [Google Scholar]
  40. Venu, M.; Venkateswarlu, S.; Reddy, Y.V.M.; Seshadri Reddy, A.; Gupta, V.K.; Yoon, M.; Madhavi, G. Highly Sensitive Electrochemical Sensor for Anticancer Drug by a Zirconia Nanoparticle-Decorated Reduced Graphene Oxide Nanocomposite. ACS Omega 2018, 3, 14597–14605. [Google Scholar] [CrossRef]
  41. Shaikshavali, P.; Madhusudana Reddy, T.; Venu Gopal, T.; Venkataprasad, G.; Kotakadi, V.S.; Palakollu, V.N.; Karpoormath, R. A simple sonochemical assisted synthesis of nanocomposite (ZnO/MWCNTs) for electrochemical sensing of Epinephrine in human serum and pharmaceutical formulation. Colloids Surf. A Physicochem. Eng. Asp. 2020, 584, 124038. [Google Scholar] [CrossRef]
  42. Jarrar, Q.B.; Hakim, M.N.; Zakaria, Z.A.; Cheema, M.S.; Moshawih, S. Renal ultrastructural alterations induced by various preparations of mefenamic acid. Ultrastruct. Pathol. 2020, 44, 130–140. [Google Scholar] [CrossRef]
  43. Aslan, M.; Kirimlioglu, E.; Afsar, E.; Ceker, T.; Yilmaz, C. Increased PUFA levels in kidney epithelial cells in the course of diclofenac toxicity. Toxicol. Vitr. 2020, 66, 104836. [Google Scholar] [CrossRef]
  44. Moling, O.; Cairon, E.; Rimenti, G.; Rizza, F.; Pristera, R.; Mian, P. Severe hepatotoxicity after therapeutic doses of acetaminophen. Clin. Ther. 2006, 28, 755–760. [Google Scholar] [CrossRef]
  45. Mazer, M.; Perrone, J. Acetaminophen-Induced Nephrotoxicity: Pathophysiology, Clinical Manifestations, and Management. J. Med. Toxicol. 2008, 4, 2–6. [Google Scholar] [CrossRef] [PubMed]
  46. Nicolle, E.; Devillier, P.; Delanoy, B.; Durand, C.; Bessard, G. Therapeutic monitoring of nalbuphine: Transplacental transfer and estimated pharmacokinetics in the neonate. Eur. J. Clin. Pharmacol. 1996, 49, 485–489. [Google Scholar] [CrossRef] [PubMed]
  47. Starakis, I.; Lekkou, A.; Blikas, A.; Labropoulou-Karatza, C. Drug-induced cardiotoxicity due to aminophylline treatment: A case report. Curr. Ther. Res. Clin. Exp. 2003, 64, 367–374. [Google Scholar] [CrossRef] [PubMed]
  48. Kiang, T.K.L.; Ranamukhaarachchi, S.A.; Ensom, M.H.H. Revolutionizing Therapeutic Drug Monitoring with the Use of Interstitial Fluid and Microneedles Technology. Pharmaceutics 2017, 9, 43. [Google Scholar] [CrossRef]
  49. Chiu, M.L.; Lawi, W.; Snyder, S.T.; Wong, P.K.; Liao, J.C.; Gau, V. Matrix Effects—A Challenge Toward Automation of Molecular Analysis. JALA Charlottesv. Va. 2010, 15, 233–242. [Google Scholar] [CrossRef]
  50. Kaczor-Urbanowicz, K.E.; Carreras-Presas, C.M.; Aro, K.; Tu, M.; Garcia-Godoy, F.; Wong, D.T.W. Saliva diagnostics—Current views and directions. Exp. Biol. Med. 2017, 242, 459–472. [Google Scholar] [CrossRef]
  51. de Castro, M.D.L. Sweat as a clinical sample: What is done and what should be done. Bioanalysis 2016, 8, 85–88. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, Z.; Liu, J.; Cheng, Y.; Chen, J.; Zhao, H.; Ren, X. Urine Analysis has a Very Broad Prospect in the Future. Front. Anal. Sci. 2022, 1, 812301. [Google Scholar] [CrossRef]
  53. Patsalos, P.N.; Berry, D.J. Therapeutic Drug Monitoring of Antiepileptic Drugs by Use of Saliva. Ther. Drug Monit. 2013, 35, 4–29. [Google Scholar] [PubMed]
  54. Patsalos, P.N.; Spencer, E.P.; Berry, D.J. Therapeutic Drug Monitoring of Antiepileptic Drugs in Epilepsy: A 2018 Update. Ther. Drug Monit. 2018, 40, 526–548. [Google Scholar] [CrossRef]
  55. Vasudev, A.; Tripathi, K.D.; Puri, V. Correlation of serum and salivary carbamazepine concentration in epileptic patients: Implications for therapeutic drug monitoring. Neurol. India. 2002, 50, 60–62. [Google Scholar]
  56. Tsiropoulos, I.; Kristensen, O.; Klitgaard, N.A. Saliva and serum concentration of lamotrigine in patients with epilepsy. Ther. Drug Monit. 2000, 22, 517–521. [Google Scholar] [CrossRef]
  57. Meearelli, O.; Voti, P.L.; Pro, S.; Romolo, F.S.; Rotolo, M.; Pulitano, P.; Accornero, N.; Vanacore, N. Saliva and serum levetiracetam concentrations in patients with epilepsy. Ther. Drug Monit. 2007, 29, 313–318. [Google Scholar] [CrossRef]
  58. Li, R.R.; Sheng, X.Y.; Ma, L.Y.; Yao, H.X.; Cai, L.X.; Chen, C.Y.; Zhu, S.N.; Zhou, Y.; Wu, Y.; Cui, Y.M. Saliva and Plasma Monohydroxycarbamazepine Concentrations in Pediatric Patients with Epilepsy. Ther. Drug Monit. 2016, 38, 365–370. [Google Scholar] [CrossRef]
  59. Chiappin, S.; Antonelli, G.; Gatti, R.; De Palo, E.F. Saliva specimen: A new laboratory tool for diagnostic and basic investigation. Clin. Chim. Acta 2007, 383, 30–40. [Google Scholar] [CrossRef]
  60. Diao, X.; Huestis, M.A. Approaches, Challenges, and Advances in Metabolism of New Synthetic Cannabinoids and Identification of Optimal Urinary Marker Metabolites. Clin. Pharmacol. Therapeutics. 2017, 101, 239–253. [Google Scholar] [CrossRef]
  61. Chung, S.; Singh, N.K.; Gribkoff, V.K.; Hall, D.A. Electrochemical Carbamazepine Aptasensor for Therapeutic Drug Monitoring at the Point of Care. ACS Omega 2022, 7, 39097–39106. [Google Scholar] [CrossRef] [PubMed]
  62. Atta, N.F.; Galal, A.; Ahmed, Y.M.; El-Ads, E.H. Design strategy and preparation of a conductive layered electrochemical sensor for simultaneous determination of ascorbic acid, dobutamine, acetaminophen and amlodipine. Sens. Actuators B Chem. 2019, 297, 126648. [Google Scholar] [CrossRef]
  63. Hanssen, B.L.; Siraj, S.; Wong, D.K.Y. Recent strategies to minimise fouling in electrochemical detection systems. Rev. Anal. Chem. 2016, 35, 1–28. [Google Scholar] [CrossRef]
  64. Ibrahim, M.; Ibrahim, H.; Almandil, N.; Kawde, A.-N. Gold nanoparticles/f-MWCNT nanocomposites modified glassy carbon paste electrode as a novel voltammetric sensor for the determination of cyproterone acetate in pharmaceutical and human body fluids. Sens. Actuators B Chem. 2018, 274, 123–132. [Google Scholar] [CrossRef]
  65. Shalauddin, M.; Akhter, S.; Basirun, W.J.; Bagheri, S.; Anuar, N.S.; Johan, M.R. Hybrid nanocellulose/f-MWCNTs nanocomposite for the electrochemical sensing of diclofenac sodium in pharmaceutical drugs and biological fluids. Electrochim. Acta 2019, 304, 323–333. [Google Scholar] [CrossRef]
  66. Atta, N.F.; Galal, A.; Ahmed, Y.M. Electrochemical Method for the Determination of Three New Anti-Hepatitis C Drugs: Application in Human Blood Serum. J. Electrochem. Soc. 2018, 165, B442. [Google Scholar] [CrossRef]
  67. Alavi-Tabari, S.A.R.; Khalilzadeh, M.A.; Karimi-Maleh, H. Simultaneous determination of doxorubicin and dasatinib as two breast anticancer drugs uses an amplified sensor with ionic liquid and ZnO nanoparticle. J. Electroanal. Chem. 2018, 811, 84–88. [Google Scholar] [CrossRef]
  68. Veera Manohara Reddy, Y.; Sravani, B.; Maseed, H.; Łuczak, T.; Osińska, M.; SubramanyamSarma, L.; Srikanth, V.V.S.S.; Madhavi, G. Ultrafine Pt–Ni bimetallic nanoparticles anchored on reduced graphene oxide nanocomposites for boosting electrochemical detection of dopamine in biological samples. New J. Chem. 2018, 42, 16891–16901. [Google Scholar] [CrossRef]
  69. Murugan, E.; Kumar, K. Fabrication of SnS/TiO2@GO Composite Coated Glassy Carbon Electrode for Concomitant Determination of Paracetamol, Tryptophan, and Caffeine in Pharmaceutical Formulations. Anal. Chem. 2019, 91, 5667–5676. [Google Scholar] [CrossRef]
  70. Kalambate, P.K.; Li, Y.; Shen, Y.; Huang, Y. Mesoporous Pd@Pt core–shell nanoparticles supported on multi-walled carbon nanotubes as a sensing platform: Application in simultaneous electrochemical detection of anticancer drugs doxorubicin and dasatinib. Anal. Methods 2019, 11, 443–453. [Google Scholar] [CrossRef]
  71. Farahani, K.Z.; Benvidi, A.; Rezaeinasab, M.; Abbasi, S.; Abdollahi-Alibeik, M.; Rezaeipoor-Anari, A.; Zarchi, M.A.K.; Abadi, S.S.A.D.M. Potentiality of PARAFAC approaches for simultaneous determination of N-acetylcysteine and acetaminophen based on the second-order data obtained from differential pulse voltammetry. Talanta 2019, 192, 439–447. [Google Scholar] [CrossRef] [PubMed]
  72. Srivastava, M.; Tiwari, P.; Mall, V.K.; Srivastava, S.K.; Prakash, R. Voltammetric determination of the antimalarial drug chloroquine using a glassy carbon electrode modified with reduced graphene oxide on WS2 quantum dots. Microchim. Acta. 2019, 186, 415. [Google Scholar] [CrossRef] [PubMed]
  73. Stoian, I.-A.; Iacob, B.-C.; Dudaș, C.-L.; Barbu-Tudoran, L.; Bogdan, D.; Marian, I.O.; Bodoki, E.; Oprean, R. Biomimetic electrochemical sensor for the highly selective detection of azithromycin in biological samples. Biosens. Bioelectron. 2020, 155, 112098. [Google Scholar] [CrossRef] [PubMed]
  74. Jandaghi, N.; Jahani, S.; Foroughi, M.M.; Kazemipour, M.; Ansari, M. Cerium-doped flower-shaped ZnO nano-crystallites as a sensing component for simultaneous electrochemical determination of epirubicin and methotrexate. Microchim. Acta 2019, 187, 24. [Google Scholar] [CrossRef]
  75. Veera Manohara Reddy, Y.; Sravani, B.; Łuczak, T.; Mallikarjuna, K.; Madhavi, G. An ultra-sensitive rifampicin electrochemical sensor based on titanium nanoparticles (TiO2) anchored reduced graphene oxide modified glassy carbon electrode. Colloids Surf. A Physicochem. Eng. Asp. 2021, 608, 125533. [Google Scholar] [CrossRef]
  76. Abdel-Haleem, F.M.; Gamal, E.; Rizk, M.S.; Madbouly, A.; El Nashar, R.M.; Anis, B.; Elnabawy, H.M.; Khalil, A.S.G.; Barhoum, A. Molecularly Imprinted Electrochemical Sensor-Based Fe2O3@MWCNTs for Ivabradine Drug Determination in Pharmaceutical Formulation, Serum, and Urine Samples. Front. Bioeng. Biotechnol. 2021, 9, 648704. [Google Scholar] [CrossRef]
  77. Monsef, R.; Salavati-Niasari, M. Hydrothermal architecture of Cu5V2O10 nanostructures as new electro-sensing catalysts for voltammetric quantification of mefenamic acid in pharmaceuticals and biological samples. Biosens. Bioelectron. 2021, 178, 113017. [Google Scholar] [CrossRef]
  78. Valian, M.; Khoobi, A.; Salavati-Niasari, M. Synthesis, characterization and electrochemical sensors application of Tb2Ti2O7 nanoparticle modified carbon paste electrode for the sensing of mefenamic acid drug in biological samples and pharmaceutical industry wastewater. Talanta 2022, 247, 123593. [Google Scholar] [CrossRef]
  79. Wentland, L.; Downs, C.; Fu, E. Comparison of signal enhancement strategies for carbamazepine detection in undiluted human saliva using an electrochemical sensor with stencil-printed carbon electrodes. Anal. Methods 2022, 14, 3103–3114. [Google Scholar] [CrossRef]
  80. Wentland, L.; Cook, J.M.; Minzlaff, J.; Ramsey, S.A.; Johnston, M.L.; Fu, E. Field-use device for the electrochemical quantification of carbamazepine levels in a background of human saliva. J. Appl. Electrochem. 2023, 53, 523–534. [Google Scholar] [CrossRef]
  81. Gomes, N.O.; Raymundo-Pereira, P.A. On-Site Therapeutic Drug Monitoring of Paracetamol Analgesic in Non-Invasively Collected Saliva for Personalized Medicine. Small 2023, 19, e2206753. [Google Scholar] [CrossRef]
  82. Lin, S.Y.; Yu, W.Z.; Wang, B.; Zhao, Y.C.; En, K.; Zhu, J.L.; Cheng, X.; Zhou, C.; Lin, H.; Wang, Z.; et al. Noninvasive wearable electroactive pharmaceutical monitoring for personalized therapeutics. Proc. Natl. Acad. Sci. USA 2020, 117, 19017–19025. [Google Scholar]
  83. Tai, L.C.; Liaw, T.S.; Lin, Y.J.; Nyein, H.Y.Y.; Bariya, M.; Ji, W.B.; Hettick, M.; Zhao, C.; Zhao, J.; Hou, L.; et al. Wearable Sweat Band for Noninvasive Levodopa Monitoring. Nano Lett. 2019, 19, 6346–6351. [Google Scholar] [CrossRef] [PubMed]
  84. Nyein, H.Y.Y.; Bariya, M.; Tran, B.; Ahn, C.H.; Brown, B.J.; Ji, W.B.; Davis, N.; Javey, A. A wearable patch for continuous analysis of thermoregulatory sweat at rest. Nat. Commun. 2021, 12, 1823. [Google Scholar] [CrossRef] [PubMed]
  85. Moon, J.M.; Teymourian, H.; De la Paz, E.; Sempionatto, J.R.; Mahato, K.; Sonsa-ard, T.; Huang, N.; Longardner, K.; Litvan, I.; Wang, J. Non-Invasive Sweat-Based Tracking of L-Dopa Pharmacokinetic Profiles Following an Oral Tablet Administration. Angew. Chem. Int. Ed. 2021, 60, 19074–19078. [Google Scholar] [CrossRef] [PubMed]
  86. Raymundo-Pereira, P.A.; Gomes, N.O.; Machado, S.A.S.; Oliveira, O.N., Jr. Wearable glove-embedded sensors for therapeutic drug monitoring in sweat for personalized medicine. Chem. Eng. J. 2022, 435, 135047. [Google Scholar] [CrossRef]
  87. Windmiller, J.R.; Zhou, N.; Chuang, M.-C.; Valdés-Ramírez, G.; Santhosh, P.; Miller, P.R.; Narayan, R.; Wang, J. Microneedle array-based carbon paste amperometric sensors and biosensors. Analyst 2011, 136, 1846–1851. [Google Scholar] [CrossRef]
  88. Goud, K.Y.; Moonla, C.; Mishra, R.K.; Yu, C.; Narayan, R.; Litvan, I.; Wang, J. Wearable Electrochemical Microneedle Sensor for Continuous Monitoring of Levodopa: Toward Parkinson Management. ACS Sens. 2019, 4, 2196–2204. [Google Scholar] [CrossRef]
  89. Goud, K.Y.; Mahato, K.; Teymourian, H.; Longardner, K.; Litvan, I.; Wang, J. Wearable electrochemical microneedle sensing platform for real-time continuous interstitial fluid monitoring of apomorphine: Toward Parkinson management. Sens. Actuators B Chem. 2022, 354, 131234. [Google Scholar] [CrossRef]
  90. Lin, S.; Cheng, X.; Zhu, J.; Wang, B.; Jelinek, D.; Zhao, Y.; Wu, T.-Y.; Horrillo, A.; Tan, J.; Yeung, J.; et al. Wearable microneedle-based electrochemical aptamer biosensing for precision dosing of drugs with narrow therapeutic windows. Sci. Adv. 2022, 8, eabq4539. [Google Scholar] [CrossRef]
  91. Ishii, K.; Ogata, G.; Einaga, Y. Electrochemical detection of triamterene in human urine using boron-doped diamond electrodes. Biosens. Bioelectron. 2022, 217, 114666. [Google Scholar] [CrossRef] [PubMed]
  92. Kimuam, K.; Rodthongkum, N.; Ngamrojanavanich, N.; Chailapakul, O.; Ruecha, N. Single step preparation of platinum nanoflowers/reduced graphene oxide electrode as a novel platform for diclofenac sensor. Microchem. J. 2020, 155, 104744. [Google Scholar] [CrossRef]
  93. Tajik, S.; Beitollahi, H.; Shahsavari, S.; Nejad, F.G. Simultaneous and selective electrochemical sensing of methotrexate and folic acid in biological fluids and pharmaceutical samples using Fe3O4/ppy/Pd nanocomposite modified screen printed graphite electrode. Chemosphere 2022, 291, 132736. [Google Scholar] [CrossRef] [PubMed]
  94. Hassan, S.S.M.; Kamel, A.H.; Fathy, M.A. A novel screen-printed potentiometric electrode with carbon nanotubes/polyaniline transducer and molecularly imprinted polymer for the determination of nalbuphine in pharmaceuticals and biological fluids. Anal. Chim. Acta 2022, 1227, 340239. [Google Scholar] [CrossRef]
  95. Liu, P.; Zhang, R.; Zheng, L.; Cao, Q. An Electrochemical Sensor for Trimethoprim Based on a Magnetic Molecularly Imprinted Carbon Paste Electrode. Chemosensors 2023, 11, 339. [Google Scholar] [CrossRef]
  96. Saher, A.; Bahgat, A.; Molouk, A.; Mortada, W.; Khalifa, M. MIP/GO/GCE Sensor for the Determination of Aminophylline in Pharmaceutical Ingredients and Urine Samples. Anal. Bioanal. Chem. Research. 2023, 10, 435–443. [Google Scholar]
  97. Sánchez-Tirado, E.; González-Cortés, A.; Yáñez-Sedeño, P.; Pingarrón, J.M. Electrochemical immunosensor for the determination of the cytokine interferon gamma (IFN-γ) in saliva. Talanta 2020, 211, 120761. [Google Scholar] [CrossRef]
  98. Beduk, D.; Beduk, T.; de Oliveira Filho, J.I.; Ait Lahcen, A.; Aldemir, E.; Guler Celik, E.; Salama, K.N.; Timur, S. Smart Multiplex Point-of-Care Platform for Simultaneous Drug Monitoring. ACS Appl. Mater. Interfaces 2023, 15, 37247–37258. [Google Scholar] [CrossRef]
  99. Lisboa, T.P.; Alves, G.F.; de Faria, L.V.; de Souza, C.C.; Matos, M.A.C.; Matos, R.C. 3D-printed electrode an affordable sensor for sulfanilamide monitoring in breast milk, synthetic urine, and pharmaceutical formulation samples. Talanta 2022, 247, 123610. [Google Scholar] [CrossRef]
  100. Tajik, S.; Sharifi, F.; Aflatoonian, B.; Di Bartolomeo, A. A New Electrochemical Sensor for the Detection of Ketoconazole Using Carbon Paste Electrode Modified with Sheaf-like Ce-BTC MOF Nanostructure and Ionic Liquid. Nanomaterials 2023, 13, 523. [Google Scholar] [CrossRef]
  101. Tachi, T.; Kaji, N.; Tokeshi, M.; Baba, Y. Simultaneous Separation, Metering, and Dilution of Plasma from Human Whole Blood in a Microfluidic System. Anal. Chem. 2009, 81, 3194–3198. [Google Scholar] [CrossRef] [PubMed]
  102. Helton, K.L.; Nelson, K.E.; Fu, E.; Yager, P. Conditioning saliva for use in a microfluidic biosensor. Lab. Chip. 2008, 8, 1847–1851. [Google Scholar] [CrossRef] [PubMed]
  103. Kava, A.A.; Beardsley, C.; Hofstetter, J.; Henry, C.S. Disposable glassy carbon stencil printed electrodes for trace detection of cadmium and lead. Anal. Chim. Acta 2020, 1103, 58–66. [Google Scholar] [CrossRef]
  104. Chen, Z.; Patel, R.; Berry, J.; Keyes, C.; Satterfield, C.; Simmons, C.; Neeson, A.; Cao, X.; Wu, Q. Development of Screen-Printable Nafion Dispersion for Electrochemical Sensor. Appl. Sci. 2022, 12, 6533. [Google Scholar] [CrossRef]
Figure 1. Traditional therapeutic drug monitoring requires clinic-based sample collection and lab-based analyses such that its utility is severely limited (outer cycle). The use of noninvasive biological fluids for sampling and field-use sensor systems for analysis could enable improved personalized therapeutic drug monitoring for each individual (inner cycle).
Figure 1. Traditional therapeutic drug monitoring requires clinic-based sample collection and lab-based analyses such that its utility is severely limited (outer cycle). The use of noninvasive biological fluids for sampling and field-use sensor systems for analysis could enable improved personalized therapeutic drug monitoring for each individual (inner cycle).
Chemosensors 11 00467 g001
Figure 2. Strategies for signal-enhanced drug monitoring using modifications to the base carbon electrodes. (A) Ibrahim et al. [64] demonstrated the quantitative detection of cyproterone acetate using glassy carbon electrodes modified with multi-walled carbon nanotubes and gold nanoparticles (left). This strategy resulted in a substantial increase in the current peak of the square-wave voltammogram relative to the case of the base electrode only (right). The voltammograms represent cyproterone acetate (4.2 µM) acquired using electrodes containing 0% (1), 0.5% (2), 1.5% (3), 2.5% (4), and 4% (5) of their nanocomposite modification (as shown in the inset plot). Reproduced with permission from Elsevier. (B) Shalauddin et al. [65] used a combination of nanocellulose and multi-walled carbon nanotubes (SEM image, left) to modify their base glassy carbon electrodes and demonstrated an improved and enhanced signal of diclofenac sodium using cyclic voltammetry (right). Reproduced with permission from Elsevier. (C) Atta et al. [62] used a composite of carbon nanotubes, ionic liquid crystals, reduced graphene oxide, and 18-Crown-6 to modify their glassy carbon electrode (left). They demonstrated the simultaneous detection of amlodipine, acetaminophen, dobutamine, and ascorbic acid using differential pulse voltammetry (right, analyte peaks from right to left); the voltammograms represent increasing concentrations of analytes in the ranges, 0.02 to 20 µM, 0.004 to 10 µM, 0.02 to 15 µM, and 0.4 to 40 µM, respectively. Reproduced with permission from Elsevier.
Figure 2. Strategies for signal-enhanced drug monitoring using modifications to the base carbon electrodes. (A) Ibrahim et al. [64] demonstrated the quantitative detection of cyproterone acetate using glassy carbon electrodes modified with multi-walled carbon nanotubes and gold nanoparticles (left). This strategy resulted in a substantial increase in the current peak of the square-wave voltammogram relative to the case of the base electrode only (right). The voltammograms represent cyproterone acetate (4.2 µM) acquired using electrodes containing 0% (1), 0.5% (2), 1.5% (3), 2.5% (4), and 4% (5) of their nanocomposite modification (as shown in the inset plot). Reproduced with permission from Elsevier. (B) Shalauddin et al. [65] used a combination of nanocellulose and multi-walled carbon nanotubes (SEM image, left) to modify their base glassy carbon electrodes and demonstrated an improved and enhanced signal of diclofenac sodium using cyclic voltammetry (right). Reproduced with permission from Elsevier. (C) Atta et al. [62] used a composite of carbon nanotubes, ionic liquid crystals, reduced graphene oxide, and 18-Crown-6 to modify their glassy carbon electrode (left). They demonstrated the simultaneous detection of amlodipine, acetaminophen, dobutamine, and ascorbic acid using differential pulse voltammetry (right, analyte peaks from right to left); the voltammograms represent increasing concentrations of analytes in the ranges, 0.02 to 20 µM, 0.004 to 10 µM, 0.02 to 15 µM, and 0.4 to 40 µM, respectively. Reproduced with permission from Elsevier.
Chemosensors 11 00467 g002
Figure 3. Advances in monitoring at the point of care. (A) Chung et al. [61] demonstrated the sensing of carbamazepine using a (diluted) fingerstick blood sample (top). In their electrochemical sensor, aptamer binding to carbamazepine resulted in an increased current signal from colocalization of the methylene blue tag with the electrode and enabled carbamazepine quantification down to 10 nM (bottom). Reproduced from [61]. (B) Wentland et al. [79,80] used the anionic surfactant sodium dodecyl sulfate (SDS) to facilitate the electrochemical detection of carbamazepine against a background of commercially purchased saliva. Their field-use compatible flow cell that included a dry SDS film on stencil-printed electrodes (left and right top) produced a similarly enhanced signal to that produced from SDS in solution (right bottom). Reproduced from [79,80]. (C) Gomes et al. [81] demonstrated acetaminophen monitoring from the unprocessed saliva of a healthy subject (top). Their sensor, pretreated using cyclic voltammetry of a sulfuric acid solution, enabled the tracking of the drug concentration over time (bottom). Reproduced from [81].
Figure 3. Advances in monitoring at the point of care. (A) Chung et al. [61] demonstrated the sensing of carbamazepine using a (diluted) fingerstick blood sample (top). In their electrochemical sensor, aptamer binding to carbamazepine resulted in an increased current signal from colocalization of the methylene blue tag with the electrode and enabled carbamazepine quantification down to 10 nM (bottom). Reproduced from [61]. (B) Wentland et al. [79,80] used the anionic surfactant sodium dodecyl sulfate (SDS) to facilitate the electrochemical detection of carbamazepine against a background of commercially purchased saliva. Their field-use compatible flow cell that included a dry SDS film on stencil-printed electrodes (left and right top) produced a similarly enhanced signal to that produced from SDS in solution (right bottom). Reproduced from [79,80]. (C) Gomes et al. [81] demonstrated acetaminophen monitoring from the unprocessed saliva of a healthy subject (top). Their sensor, pretreated using cyclic voltammetry of a sulfuric acid solution, enabled the tracking of the drug concentration over time (bottom). Reproduced from [81].
Chemosensors 11 00467 g003
Figure 4. Advances in the electrochemical monitoring of drug analytes in sweat. (A) Lin et al. [82] developed a sensor for acetaminophen in sweat (left). A highlight of their design was the surface engineering of their boron-doped diamond electrode with a passivation layer combined with a Nafion coating to increase the peak current as well as to shift the peak potential for acetaminophen away from the potentials of interferents (right). In the schematic plot, orange corresponds to the analyte of interest, dashed grey to interferents, and blue to the overall signal. Reproduced from [82]. (B) Tai et al. [83] reported the development of an arm band sensor for levodopa monitoring in stimulated sweat. Their sensor design uses gold electrodes on a plastic substrate to measure the tyrosinase-catalyzed oxidation of levodopa (top). The sensor showed an increasing current signal for greater concentrations of levodopa against a background of sweat (bottom). Reprinted with permission from [83]. Copyright 2019 American Chemical Society. (C) Nyein et al. [84] developed a wearable patch that enables the monitoring of levodopa in sweat at rest. Their multilayer design also includes a hydrogel layer that is coupled to a hydrophilic spacer for the transport of small volumes of sweat from the skin to the downstream microfluidic channel for analysis (left). Their patch can simultaneously measure heart rate, sweat rate, and sweat levodopa levels (also via tyrosinase-catalyzed oxidation) in situ as demonstrated in a healthy human subject after consuming L-dopa containing broad beans (right). Reproduced from [84].
Figure 4. Advances in the electrochemical monitoring of drug analytes in sweat. (A) Lin et al. [82] developed a sensor for acetaminophen in sweat (left). A highlight of their design was the surface engineering of their boron-doped diamond electrode with a passivation layer combined with a Nafion coating to increase the peak current as well as to shift the peak potential for acetaminophen away from the potentials of interferents (right). In the schematic plot, orange corresponds to the analyte of interest, dashed grey to interferents, and blue to the overall signal. Reproduced from [82]. (B) Tai et al. [83] reported the development of an arm band sensor for levodopa monitoring in stimulated sweat. Their sensor design uses gold electrodes on a plastic substrate to measure the tyrosinase-catalyzed oxidation of levodopa (top). The sensor showed an increasing current signal for greater concentrations of levodopa against a background of sweat (bottom). Reprinted with permission from [83]. Copyright 2019 American Chemical Society. (C) Nyein et al. [84] developed a wearable patch that enables the monitoring of levodopa in sweat at rest. Their multilayer design also includes a hydrogel layer that is coupled to a hydrophilic spacer for the transport of small volumes of sweat from the skin to the downstream microfluidic channel for analysis (left). Their patch can simultaneously measure heart rate, sweat rate, and sweat levodopa levels (also via tyrosinase-catalyzed oxidation) in situ as demonstrated in a healthy human subject after consuming L-dopa containing broad beans (right). Reproduced from [84].
Chemosensors 11 00467 g004
Figure 5. Advances in electrode integration into microneedles for the electrochemical sensing of drug analytes in ISF. (A) Goud et al. [89] developed an electrochemical sensor for the monitoring of apomorphine in ISF. A highlight of their sensor design is the integration of modified carbon paste electrodes into 3D printed microneedles. The electrodes were coated with a 1% Nafion film to minimize fouling from protein interferents. The addition of 2% rhodium nanoparticles was demonstrated to increase the signal generated for the oxidation of apomorphine. Their sensor showed stability across multiple measurements (10 with intervals of 10 min) and quantitative response across concentrations from 0 to 60 µM. Reproduced from [89] with permission from Elsevier. (B) Lin et al. [90] developed an electrochemical sensor for monitoring antibiotics in ISF. Their design uses electrodes composed of gold-coated acupuncture needles. The sensing mechanism used aptamers tagged with methylene blue, such that aptameter binding to the antibiotic results in the positioning of the methylene blue in closer proximity to the electrode and an increase in the current signal from the methylene blue activity. Their sensor showed quantitative response across vancomycin concentrations from 0 to 100 µM (arrow denotes increasing concentrations for the series of voltammograms). Reproduced from [90].
Figure 5. Advances in electrode integration into microneedles for the electrochemical sensing of drug analytes in ISF. (A) Goud et al. [89] developed an electrochemical sensor for the monitoring of apomorphine in ISF. A highlight of their sensor design is the integration of modified carbon paste electrodes into 3D printed microneedles. The electrodes were coated with a 1% Nafion film to minimize fouling from protein interferents. The addition of 2% rhodium nanoparticles was demonstrated to increase the signal generated for the oxidation of apomorphine. Their sensor showed stability across multiple measurements (10 with intervals of 10 min) and quantitative response across concentrations from 0 to 60 µM. Reproduced from [89] with permission from Elsevier. (B) Lin et al. [90] developed an electrochemical sensor for monitoring antibiotics in ISF. Their design uses electrodes composed of gold-coated acupuncture needles. The sensing mechanism used aptamers tagged with methylene blue, such that aptameter binding to the antibiotic results in the positioning of the methylene blue in closer proximity to the electrode and an increase in the current signal from the methylene blue activity. Their sensor showed quantitative response across vancomycin concentrations from 0 to 100 µM (arrow denotes increasing concentrations for the series of voltammograms). Reproduced from [90].
Chemosensors 11 00467 g005
Table 1. Classes of pharmaceuticals for which drug monitoring could be beneficial.
Table 1. Classes of pharmaceuticals for which drug monitoring could be beneficial.
ClassExample(s) from the Studies ReviewedMotivation for Drug Monitoring
Drugs to suppress seizures in the treatment of epilepsyCarbamazepineHighly variable pharmacokinetics; strong interactions with other common drugs; and/or high toxicity [13,14]
Drugs to treat bacterial or fungal infectionsTobramycin; vancomycin; levofloxacin; rifampicin; azithromycin; sulfanilamide; trimethoprim; ketoconazole Potential for kidney injury for tobramycin [29]
Drugs to treat Parkinson’s diseaseApomorphine; levodopaSide effects of nausea for apomorphine [30] and decreased efficacy and increased motor disturbances with use for levodopa [31]
Drugs to treat depressionParoxetine; benzodiazepines Highly variable pharmacokinetics with a longer time for clearance with aging and renal/hepatic damage for paroxetine [26]
Drugs to treat cancerMethotrexate; doxorubicin; dasatinib; epirubicin; cyproterone acetate; regorafenib; interferon gammaPulmonary and hepatotoxicity for methotrexate [32]; cardiotoxicity for doxorubicin [33]; adverse effects include dyspnea, fatigue, nausea for dasatinib [34]; hepatotoxicity for epirubicin [35]; hepatotoxicity for cyproterone acetate [36]; adverse effects include dyspnea, fatigue, nausea for regorafenib [37]
Drugs to treat hepatitis C viral infectionDaclatasvir; sofosbuvir; ledipasvirPotential adverse drug–drug interactions for transplant and HIV patients [38]
Drugs to treat psychiatric disordersOlanzapineSide effects of overdose such as nausea, slurred speech, vomiting, damage to the aorta resulting in bleeding or death [39]
Drugs to treat cardiac conditionsEtilefrine; epinephrineOverdosing on etilefrine can cause heart failure, hypertension, and erectile dysfunction [40]; epinephrine has interactions with other common compounds [41]
Anti-inflammatory and analgesicMefenamic acid; diclofenac sodiumPotential for renal toxicity [42,43]
AnalgesicAcetaminophen/
paracetamol; nalbuphine
Hepatotoxicity [44] and nephrotoxicity [45] for acetaminophen; potential for TDM in neonates for nalbuphine [46]
Drugs for the treatment of bronchial asthmaAminophyllinePotential for drug-induced cardiotoxicity [47]
Table 2. Summary of the composition and challenges of drug analyte detection in complex matrices.
Table 2. Summary of the composition and challenges of drug analyte detection in complex matrices.
Complex BiofluidMajor ComponentsAdvantagesDisadvantages
BloodIons, proteins, glucose, amino acids, lipids, hormones, erythrocytes, leukocytes, platelets [49]Gold standard; uniform across individuals; small fingerstick volumes (20 µL) are compatible with point-of-care collectionInvasive and painful; larger venipuncture volumes require a phlebotomist, which is inconvenient and limited to low-frequency collection
SerumIons, proteins, glucose, amino acids, lipids, hormones [49]
SalivaIons, small molecules, proteins, mucins, hormones, blood-derived compounds, food debris, uric acid [59]Noninvasive; moderate (1 mL) volume; easy to sample; frequent donation possible and on demand; could be compatible with continuous wearable deviceProperties variable across individuals; variability throughout the day for individuals including pH; possible food contamination
SweatIons, small molecules, proteins, pyruvate, lactate urea, antigens, antibodies, ethanol [51]Noninvasive; compatible with continuous wearable devicesLow secretion rate (10 to 100 nL/min per cm2) volume unless stimulated; variability in the rate secreted; possible contamination from cosmetics or the environment
UrineInorganic salts, urea, uric acid, proteins, enzymes, nucleic acids, vitamins, hormones, amino acids, mesothelin, beta-microglobulin, antibiotics, urokinase, mycomycin [52]Noninvasive; large (many mL) volume; easy to sample; there may be a longer time window available for drug detection compared to other biological matrices such as saliva or blood [60]Sampling is not always possible ‘on demand’; contamination potential if the collection is not conducted carefully [60]
Interstitial fluidAmino acids, carbohydrates, fatty acids [48]Noninvasive; compatible with continuous wearable devicesVery small (nL) volume unless suction used
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

Fu, E.; Khederlou, K.; Lefevre, N.; Ramsey, S.A.; Johnston, M.L.; Wentland, L. Progress on Electrochemical Sensing of Pharmaceutical Drugs in Complex Biofluids. Chemosensors 2023, 11, 467. https://doi.org/10.3390/chemosensors11080467

AMA Style

Fu E, Khederlou K, Lefevre N, Ramsey SA, Johnston ML, Wentland L. Progress on Electrochemical Sensing of Pharmaceutical Drugs in Complex Biofluids. Chemosensors. 2023; 11(8):467. https://doi.org/10.3390/chemosensors11080467

Chicago/Turabian Style

Fu, Elain, Khadijeh Khederlou, Noël Lefevre, Stephen A. Ramsey, Matthew L. Johnston, and Lael Wentland. 2023. "Progress on Electrochemical Sensing of Pharmaceutical Drugs in Complex Biofluids" Chemosensors 11, no. 8: 467. https://doi.org/10.3390/chemosensors11080467

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