detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through...

11

Click here to load reader

Upload: martin-bezdomny

Post on 11-May-2015

388 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

ORIGINAL PAPER

Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

Anupama Aggarwal & Mengjia Hu & Ingrid Fritsch

Received: 31 August 2012 /Revised: 2 December 2012 /Accepted: 12 January 2013 /Published online: 9 February 2013# Springer-Verlag Berlin Heidelberg 2013

Abstract The electrochemical behavior of dopamine wasexamined under redox cycling conditions in the presenceand absence of a high concentration of the interferent ascorbicacid at a coplanar, microelectrode array where the area of thegenerator electrodes was larger than that of the collectorelectrodes. Redox cycling converts a redox species betweenits oxidized and reduced forms by application of suitablepotentials on a set of closely located generator and collectorelectrodes. It allows signal amplification and discriminationbetween species that undergo reversible and irreversible elec-tron transfer. Microfabrication was used to produce 18 indi-vidually addressable, 4-μm-wide gold band electrodes, 2 mmlong, contained in an array having an interelectrode spacing of4 μm. Because the array electrodes are individually address-able, each can be selectively biased to produce an overalloptimal electrochemical response. Four adjacent microbandswere shorted together to serve as the collector, and wereflanked on each side by seven microbands shorted as thegenerator (a ratio of 1:3.5 of electroactive area, respectively).This configuration achieved a detection limit of 0.454±0.026 μM dopamine at the collector in the presence of100 μM ascorbic acid in artificial cerebrospinal fluid buffer,concentrations that are consistent with physiological levels.Enhancement by surface modification of the microelectrodearray to achieve this detection limit was unnecessary. Theresults suggest that the redox cycling method may be suitablefor in vivo quantification of transients and basal levels ofdopamine in the brain without background subtraction.

Keywords Dopamine . Ascorbic acid . Electrochemistry .

Redox cycling . Microelectrode array

Introduction

The catecholamine dopamine (DA) is an important neuro-transmitter in the brain.Dopaminergenic systems are involvedin a multitude of functions, including neurocognition, motorcontrol, motivation, punishment, and reward [1–8]. Dysfunc-tion of DA production and uptake rates plays a vital role invarious illnesses, such as Parkinson’s disease [9, 10],Huntington’s disease [11], substance abuse [10, 12, 13],Tourette’s syndrome [14], schizophrenia [15], and atten-tion deficit–hyperactivity disorders [12, 16]. The need tostudy DA for neurological and medical purposes is welldocumented in the literature.

Facile oxidation of DA to dopamine quinone (DAQ)renders it suitable for electroanalysis. Electrochemical de-tection has been used successfully to investigate DA directlyin the brain to determine its implications in processingmotor, motivational, and sensory information [17], anticipa-tory learning [18], the effects of addictive drugs such ascocaine [19] and nicotine [20], and general function [21].However, there remains a need to differentiate between theelectrochemical signal arising from DA and that arisingfrom ascorbic acid (AA), as well as other electrochemicallyinterfering species, including other catecholamines. This isbecause the oxidation of AA occurs at approximately+0.2 V versus Ag/AgCl (4 M NaCl), which is similar to thepotential for DA and other catecholamines [22, 23]. Also, itis a challenge to measure the low concentrations of DA in thepresence of high concentrations of AA and other reactivespecies. Extracellular DA concentrations from evoked re-lease in the striatum, for example, are reported to be single-digit micromolar concentrations [21, 24–26], whereas the

Published in the topical collection Bioelectroanalysis with guesteditors Nicolas Plumeré, Magdalena Gebala, and WolfgangSchuhmann.

A. Aggarwal :M. Hu : I. Fritsch (*)Department of Chemistry and Biochemistry,University of Arkansas, Fayetteville, AR 72701, USAe-mail: [email protected]

Anal Bioanal Chem (2013) 405:3859–3869DOI 10.1007/s00216-013-6738-z

Page 2: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

concentrations of AA (100–500 μM) are 100 times to 1,000times those of DA [10, 22, 23, 27–30].

Several electrochemical methods have been used formeasuring DA in vivo. Much of the earlier work employeddifferential pulse voltammetry. It has poor temporal resolu-tion (more than 30 s) and allows selectivity for simultaneousmeasurement of analytes with oxidation peaks only morethan 100 mV apart [31]. Chronoamperometry (CA) offerssubmillisecond temporal resolution. It has been used todetect DA and other catecholamine neurotransmitters[32–35], but has poor chemical selectivity. Fast-scan cyclicvoltammetry (CV) [23, 36–38] at carbon-fiber electrodes(CFEs) is currently the most widely used electrochemicalmethod for making DA measurements in vivo. It has hightemporal resolution (100 ms), although not as good as CA.At fast scan rates the current from AA oxidation is negligi-ble. However, the rapidly changing voltage causes a largebackground current due to double-layer charging that over-takes the faradaic current. Because the background currentin vivo is stable for only short times, fast-scan CV is typi-cally not used for measurements exceeding 1 min withoutreestablishing the background signal [23]. Thus, it is diffi-cult with fast-scan CV to measure resting or static DAconcentrations. This is where redox cycling may ultimatelyhave an advantage.

Because DA is electrochemically reversible, it can bedetected by redox cycling, where two or more closelyspaced electrodes, the generator and the collector, are heldat oxidative (anodic) potentials and reductive (cathodic)potentials, respectively. Charging current is not an issuebecause it drops to zero within nanoseconds to microsec-onds after the initial application of the potential at a micro-electrode. Upon arrival at the generator, DA undergoesoxidation. The DAQ product diffuses away, where somemolecules reach the collector across the gap with a certainefficiency and reduce back to DA; the cycle continues.Thus, current at both the generator and the collector can beused to determine the DA concentration. AA, however, alsooxidizes at the generator, adding to the measured anodiccurrent. The dehydroascorbic acid (DHAA) product thenundergoes subsequent rapid hydrolysis to a nonelectroactiveform (rate constant of 1.4×103s−1 [39]) prior to reaching thecollector, appearing silent at the cathode on the timescale ofthis experiment [40, 41]. Therefore, the redox cycling meth-od offers improved signal-to-background ratio and real-timemonitoring of DA without requiring high-speed scanningelectronics. It also provides elimination from AA interfer-ence when the current is monitored at the cathodic collector[42, 43].

Redox cycling to detect DA using the closely spacedelectrodes of a microelectrode array (MEA) has been previ-ously reported [42–49]. Low nanomolar concentrationshave been measured with redox cycling involving flow

channels or electrode modification [42, 44, 45, 47, 48,50–52]. However, none of those studies demonstrate simul-taneous detection of DA in the presence of a large excess ofAA (100 times) at physiological concentrations, which isimportant for in vivo detection, without requiring an additionalcoating to concentrate DA and exclude AA.

To perform the in vitro redox cycling studies describedhere, a MEAwas microfabricated. It is important to considerits advantages and disadvantages compared with those ofCFEs when considering it for in vivo measurements. Thestate-of-the-art fabrication procedure for CFEs is lengthy,manually intensive, and irreproducible. Also, the single sens-ing site of a CFE limits the scope of mapping DA in the brain.In an attempt to resolve that complication, Zachek et al. [53]evenly spaced and manually mounted four CFEs as an arrayinto a holder to perform simultaneous in vivo DA measure-ments by fast-scan CV in spatially distinct locations. Micro-fabrication of multiple electrodes on one substrate, in contrast,can provide greater control over electrode size, spacing, andflexibility in electrode design, and thus allows the more ad-vanced electrochemical methods, such as redox cycling, ana-lyte mapping, and multianalyte sensing, to be possible withoutextensive manual manipulation [54].

Two other issues to consider in comparing microfabricatedMEAs with CFEs are the suitability of the electrode materialand the probe size for in vivo analysis. Gold served as theelectrode material of our array. This is because metals can bemore easily microfabricated onto substrates than carbon withhigh spatial resolution [55–57]. Nonetheless, carbon electro-des patterned with relatively large features on probes for tissueinsertion have been reported [58, 59]. Metal electrodes (goldand platinum), however, generally have a smaller inert poten-tial range compared with carbon electrodes, which needs to bekept in mind in setting a voltage window for analysis. Withrespect to probe size, the insertion of CFEs (5–15-μm diam-eter) and metal electrodes of small diameters in the brainavoids ischemia and glial proliferation [60]. This is in contrastto the degeneration of tissue observed from insertion of muchlarger microdialysis probes (approximately 220 μm or greaterin diameter) [24, 26, 60–63], which affects the analyte con-centrations in the sampled fluid [21, 25, 64]. Thus, it isdesirable to construct microfabricated arrays to be as smallas possible, at least narrower than microdialysis probes. Al-though a systematic study of tissue damage has not yet beenperformed onmicrofabricated microelectrodes of intermediatewidth (approximately 100 μm at the sensing element), in vivoresults suggest that tissue damage is indeed less than for thelarger microdialysis probes [57, 58, 65, 66]. Similar probedimensions of 100 μm (and smaller) for the substrate arecertainly achievable for interdigitated microband arrays suchas the MEA used here for redox cycling. A next-generationredox-cycling MEA design and the effect on signal for probesto be used in vivo are discussed further in “Conclusions.”

3860 A. Aggarwal et al.

Page 3: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

To our knowledge, this is the first report of quantification ofDAwith redox cycling in the presence of up to 100 times itsconcentration of AA at physiologically relevant levels, with-out using polymer coatings that concentrate the analyte andexclude the interferent. A MEA was used with an electrodeassignment having a small central collector region flanked bytwo larger generator regions to enhance the discriminationagainst AA. To further challenge the redox cycling method,artificial cerebrospinal fluid (aCSF) was chosen as the elec-trolyte, because its composition better mimics extracellularbrain fluid [67] than the more commonly used pristine buffersused for in vitro studies, such as phosphate-buffered saline,2-[4-(2-hydroxyethyl)piperazin-1-yl]ethanesulfonic acid(HEPES)-buffered saline, and 4-amino-2-(hydroxymethyl)-propane-1,3-diol (Tris) [49, 68–70].

Experimental

Chemicals and materials

All chemicals were used as received unless otherwise stated.Hexaammineruthenium(III) chloride, dopamine hydrochloride,calcium chloride, sodium dihydrogen phosphate monohydrate,

L-(+)-AA, HEPES, and Tris (all ACS grade) were obtainedfrom Alfa Aesar (Ward Hill, MA, USA). Magnesium sulfateand sodium chloride were obtained from EMD Chemicals(Gibbstown, NJ, USA). Sodium hydrogen carbonate wasobtained from J.T. Baker (Phillipsburg, NJ, USA). Sodiumsulfate was obtained from Aldrich Chemical (St. Louis, MO,USA). D(+)-glucose (anhydrous) and potassium chloride wereobtained from BDH/VWR (Radnor, PA, USA). Water (ACSreagent grade, 18 MΩ cm or greater) was obtained fromRicca Chemical (Arlington, TX, USA).

The MEA

Details of the design of the MEA and one of the two electrodeconfigurations are shown in Fig. 1. The devices contain 18,parallel, coplanar band electrodes (expanded view in Fig. 1b)and two large gold features (4.00mm×4.00mm and 4.00mm×2.00 mm), of which the larger one served as an auxiliaryelectrode and the smaller one was not used (left at open circuit).Each of the electrodes in the array is individually addressablevia contact pads. The microbands have an average length of2.00±0.01 mm, a width of 4.0±0.1 μm, and are separated bygaps of 4.0±0.1 μm. Electrochemical characterization of theMEAs using the model compound hexaammineruthenium(III)chloride is reported elsewhere [71].

MEA fabrication is also reported in detail elsewhere [71].In brief, a 50-nm-thick layer of gold was deposited onto a5-nm-thick chromium adhesion layer on a silicon dioxidecoated (2-μm) silicon wafer. The metal layers were patterned

to define electrodes, leads, and contact pads. The MEAs werethen coated by an insulation layer of benzocyclobutene, whichwas subsequently patterned to define the active electrode areasthat would be exposed to solution and the contact pads thatwould be inserted into the edge connector (solder contact,20/40 position, 0.05-in. pitch, Sullins Electronics, San Marcos,CA, USA).

Electrochemical methods and instrumentation

All experiments were performed with a fresh solution eachtime to avoid carryover from the previous electrochemistry.A bipotentiostat equipped with a Faraday cage and apicoamp booster (CHI 760, CH Instruments, Austin, TX,USA) was used.

The electrode configuration that was used had one collectorregion, formed by shorting either two or four microbandelectrodes together. It was flanked by two larger generatorregions, each formed by shorting eight or seven microbandelectrodes together, respectively. Figure 1c illustrates the latter

- Collector

DAQ DA

(b)

(c)

DA

AA AAO

DAQ

(d) DA

AA AAO

DAQ

+ Generator + Generator Diffusion Diffusion

1

(a)

Contact Pads

Auxiliary

4 mm x 4 mm

Array(2 mm long)

50 m 2 mm

Fig. 1 Electrode chip design and assignment of microband electrodesas generators and collectors. a Photographic image showing a full viewof the microelectrode-array-containing chip. b An expanded view ofthe array region (cropped in length) showing 4-μm-wide microbandelectrodes with 4-μm gaps. c The configuration with the large gener-ator regions flanking a single, smaller collector region used for thecalibration studies. The generator electrodes are shown in black and thecollector electrodes in gray (not drawn to scale). d Redox cyclingbetween the reduced form dopamine (DA) and dopamine quinone(DAQ), the oxidized form of dopamine, and the oxidation of thereduced form of ascorbic acid (AA) to a form that is unable to re-reduceat the collector electrodes (AAO)

Detection of dopamine in the presence of excess ascorbic acid 3861

Page 4: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

configuration used to produce the calibration curves for DAobtained by CA in the presence and absence of AA.

Either CV (by sweeping from −0.100 to +0.500 V andback) or CA (at a constant oxidizing potential of +0.500 V)versus a Ag/AgCl (saturated KCl) reference electrode wasperformed at the generator electrodes. In the redox cyclingexperiments, the collector electrodes were held at a constantreducing potential (−0.100 V for CV studies and −0.200 Vfor CA studies). The potentiostat applies a potential at theelectrode(s) for 2 s (the quiet time) prior to commencing theCV and CA waveforms. (Data are not acquired during thequiet time.) Those potentials during the quiet time for CVwere −0.100 Vat the generator and −0.100 Vat the collector.Those potentials for CA were 0.000 V at the generator and−0.200 V at the collector.

Solutions containing DA and AAwere made fresh beforeuse each day in Tris buffer (15 mM Tris, 140 mM NaCl,3.25 mM KCl, 1.2 mM CaCl2, 1.2 mM MgCl2, 1.25 mMNaH2PO4, 2.0 mM Na2SO4) or in aCSF buffer (100 mMNaCl, 5.0 mM KCl, 1.2 mM NaH2PO4, 5.0 mM NaHCO3,10 mM glucose, 2.5 mM HEPES, 1.2 mM MgSO4, 1.0 mMCaCl2) at 7.4 pH. Stock solutions were purged under argongas to prevent air oxidation, but the experiments them-selves were conducted with solutions exposed to air.

The limit of detection was calculated as three times thestandard deviation of the blank signal, divided by the slopeof the calibration curve. The standard deviation of the limitof detection was obtained from propagation of error. Thecollection efficiencies, unless stated otherwise, wereobtained by taking the ratio of the steady-state current atthe collector electrode to the steady-state current at thegenerator during redox cycling. The amplification factors,unless otherwise stated, were obtained by taking the ratio ofthe current at the generator electrode undergoing redoxcycling to that at the generator electrode without redoxcycling.

Results and discussion

Characterization of 1 μM DA in 100 μM AA by CV in Trisbuffer

Redox cycling of DA (1 μM) was evaluated by slow-scanCV (0.010 Vs−1) in the presence and absence of a 100-foldexcess of AA in Tris buffer. These concentrations are withinthe ranges of those found in the brain. Figure 1d illustratesthe general redox cycling process and how DA may bedetected at the collector electrode (via reduction of DAQ)while simultaneously discriminating against AA. The redoxcycling results are shown in Fig. 2.

The procedure involved holding the potential of the collectorat a constant reducing potential of −0.100Vwhile sweeping the

potential of the generator at a scan rate of 0.010Vs−1, starting ata reducing potential of −0.100 V, where there should be no orlittle faradaic current, to +0.500 V, which is sufficient to oxidizeDA and AA, producing an anodic current. The potential at thegenerator was then scanned back to −0.100 Vat the same rate.The collector showed a cathodic current when the generatorreached a potential that was anodic enough to produce anoxidized species that diffused to the collector, where it wasreduced. The slow scan was necessary to ensure that a steadystate was achieved for the redox cycling. All of the CVresponses in Fig. 2 are plotted as a function of the potentialapplied to the generator in the forward and reverse sweeps; thecurrent in plots a, c, and e in Fig. 2 was measured at thecollector and the current in plots b, d, and f in Fig. 2 wasmeasured at the generator.

The configuration of the electrodes used for redox cyclingusing the CVapproach involved two large generator regions ofeight microbands each, flanking a much smaller collector regionof two electrodes. This configuration was chosen because it isknown that AA undergoes a spontaneous reaction with DAQ,converting it back to DA, which diminishes the amount of DAQarriving at the collector electrode, resulting in a loss of signal atthe collector, although the kinetics are slow [43, 48]. Thisarrangement, as opposed to the more popular alternating con-figuration of equally sized generator and collector electrodes,takes advantage of the shielding effect of neighboring generators[71, 72] to diminish the concentration of AA near the collector.Niwa et al. [42] described the problem of recovering DAQ at thecollector with increasing concentrations of AA at an alternatinggenerator–collector configuration of interdigitated microbandelectrodes. They showed that in the presence of a concentrationof AA that is six times that of DA, the cathodic current for DAQat the collector drops to one third (for 10-μm widths and 5-μmgaps) and two thirds (for 3-μm widths and 2-μm gaps), com-pared with the DAQ signal in the absence of AA. We wouldexpect the DAQ signal for an alternating configuration of ourarray, which has intermediate dimensions (4-μm widths and4-μm gaps), to decrease to a similar value (between one thirdand two thirds). For a solution where the concentration of AA is100 times that of DA, the decrease in the cathodic current shouldbe far worse or even immeasurable. Thus, a large generator-to-collector area was used for our studies here to consume a largeramount of AA and allow more DAQ molecules to survive thetrip to the collector from the generator.

Plot e in Fig. 2 shows an easily measured cathodiccurrent at the collector in a mixture containing 1 μM DAand 100 μM AA when the potential at the generator wassufficiently anodic (about +0.150 to +0.500 V) to produceDAQ. Its magnitude (when the generator was at +0.450 V)was 1.68 nA. This is a magnitude similar to that for 1 μMDA alone (1.55 nA; Fig. 2, plot a).

There is no contribution to the collector current from theexcess 100 μM AA because the oxidized product DHAA is

3862 A. Aggarwal et al.

Page 5: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

quickly hydrolyzed to an electrochemically irreversible form.However, as can be seen in plot c in Fig. 2, there is anoticeable signal at the collector (40 % of that when 1 μMDA is present) due to reduction of oxygen that is no longerdepleted at the generator electrodes when they are scanned topositive potentials. This explanation is further confirmed byredox cycling in buffer alone (not shown), where the collectorshowed similar behavior. This effect with oxygen providesfurther justification for exploiting large, flanking generatorelectrodes that “titrate” unwanted species (such as AA), andtherefore prevent undesired reactions with analyte speciesbefore they can be detected at the collector electrodes.

At the generator, because both DA and AA oxidize in asimilar voltage range, the CV response of the mixture shouldequal the sum of the responses for the individual components,unless there are reactions between them. Overall, the shape ofthe CVresponse for the mixture of 1 μMDA and 100 μMAA(Fig. 2, plot f) reflects the sum of the shapes for 1 μM DA(Fig. 2, plot b) and 100 μM AA (Fig. 2, plot d) alone. (Thisadditive effect at the generator is consistent with previousredox cycling results performed with DA and AA at a rela-tively high concentration of 1 mM each [49].) However, themagnitude of the generator current for the mixture at +0.450V(19.85 nA; Fig. 2, plot f) is slightly smaller than the sum of the

currents for 1 μMDA (2.92 nA in Fig. 2, plot b) and 100 μMAA (14.80 nA in Fig. 2, plot d). This suggests that there mightbe reactions between these species or their oxidized products.The fate of DA and DAQ in the presence of AA during redoxcycling is discussed in the next section in the context of dataobtained from studies using CA.

Calibration curves of DA in the presence and absenceof 100 μM AA in aCSF buffer obtained by CA

CA was used instead of CV to produce calibration curves,because although CV is more diagnostic, CA is more quan-titative and has better temporal resolution at a specifiedpotential. Ultimately, CA, not slow-scan CV, would be theelectrochemical method of choice for redox cycling for invivo measurements, where CA would allow observation ofboth transients and basal levels of DA. The response time ofthe collector is delayed from that of the generator by thetime required for DAQ to diffuse across the gap between thegenerator and the collector. That would be approximately33 ms under mass transfer conditions (based on the Einsteinequation, Eq. 1) [73],

Δ ¼ffiffiffiffiffiffiffiffi

2Dtp

; ð1Þ

Fig. 2 Comparison of cyclic-voltammetric responses at the collector(a, c, e) and the generator (b, d, f) during redox cycling at a scan rate of0.010 Vs−1 in 1 μM DA (a, b), 100 μM AA (b, c), and their mixture(d, e) in 4-amino-2-(hydroxymethyl)propane-1,3-diol buffer (pH7.4)using a configuration similar to that in Fig. 1c, but with only twomicroband electrodes serving as the collector and eight microbandelectrodes in each of the two large generator regions flanking the

smaller collector region. A triangular potential waveform was appliedto the generator, starting at −0.100 V, sweeping to +0.500 V at a scanrate of 0.010 Vs−1, and then switching direction to sweep back at thesame scan rate to +0.500 V. Meanwhile, the collector was held at aconstant reducing potential of −0.100 V. The current measured at thecollector is plotted as a function of the applied potential at the generator.All potentials are referenced to Ag/AgCl (saturated KCl)

Detection of dopamine in the presence of excess ascorbic acid 3863

Page 6: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

assuming a gap between the generator and the collector of4 μm and using a diffusion coefficient of 2.4×10−6cm2s−1

[74] for DA in striatal extracellular space.The buffer was also changed to produce the calibration

curves, from Tris to aCSF (pH7.4). The latter more closelymimics the composition of extracellular fluid in the brain,further challenging the redox cycling method and providinga more realistic representation of results that might beexpected for in vivo brain studies.

In addition, the electrode configuration was changed.Four electrodes were shorted together to form the collectorregion, instead of two, so that a larger cathodic current wouldbe obtained, and therefore improve the sensitivity for DA.The number of microbands assigned as generators was de-creased to seven on each side of the collector region, keepingthe total number of active electrodes at 18. The electrodeconfiguration is shown in Fig. 1c. This change in electrodeconfiguration was based on studies (not reported here) with awell-behaved, reversible redox species. In a solution con-taining 5 mM Ru(NH3)6Cl3 and 0.5 M KCl, the collectorcurrent increased by 19 % (from 915.1 to 1,090 nA), thegenerator current increased by 3.5 % (from 2,232 to2,311 nA), collection efficiency increased by 15 % (from41.0 % to 47.2 %), and the amplification factor increased by8.2 % (from 1.35 to 1.46). (More information on how dif-ferent assignments of collector and generator electrodes af-fect current, collection efficiencies, and amplification factorsis available elsewhere [71].)

Examples of raw CA responses under these conditionsare shown in Fig. 3. The generator current (Fig. 3, plots b, c,e, f, h, and i) responds to the potential step (from theequilibrating potential of 0.000 V to a value of +0.500 V)as expected. A large anodic faradaic transient current atinitial times (from oxidation of DA and AA) drops offapproximately with the inverse square root of time until atlong times it reaches a pseudosteady state or steady statewhen the length of the diffusion layer exceeds the generatorwidth, when mass transfer is dominated by radial diffusion,or when redox cycling becomes established, respectively.Charging current should also be superimposed on the initialfaradaic current, but it falls off exponentially with time(within microseconds) and is not distinguishable in Fig. 3.As expected, when the collector is held at −0.200 V, the DAis recycled and the anodic current at the generator is ampli-fied (Fig. 3, plots b and h), compared with the case when thecollector is left at open circuit (no redox cycling; Fig. 3,plots c and i). There is no enhancement for the solutioncontaining only AA (Fig. 3, plot e vs plot f), because thehydrolyzed DHAA cannot be recycled by the collector.

The collector current (Fig. 3, plots a, d, and g) beginsnear zero when the generator is stepped to +0.500 V. This isbecause the collector is equilibrated for 2 s prior to the stepat the same potential at which it is held during the step

(−0.200 V), which is too far negative to oxidize DA andAA, and there is no DAQ yet to reduce. Thus, the collectoris not expected to show charging or an initial cathodicfaradaic transient, with the exception of reduction of resid-ual oxygen, which is most obvious when DA is absent(Fig. 3, plot d). A cathodic current then rapidly grows atthe collector upon arrival of DAQ from the generator(Fig. 3, plots a and g), and is generated at a high rate initiallyand then more slowly as the diffusion layer at the generatoris depleted and redox cycling achieves a steady state.

Calibration curves obtained from the current measured at15 s in the CA responses for DA over a range from 5 to100 μM in the presence and absence of 100 μM AA inaCSF are shown in Fig. 4, both with and without redoxcycling at the electrode configuration illustrated in Fig. 1c.The current is linear with DA concentration (all with R2≥0.990, except for DA at the collector in the absence of AA,where R2=0.927). The y-intercept of 35 nA for the plot ofthe generator current is due to oxidation of the 100 μM AA,as expected.

There was an 8 % increase in sensitivity to DA at thegenerator electrode during redox cycling in the presence of100 μMAA (0.697±0.024 nAμM−1; Fig. 4b) compared withDA alone (0.645±0.031 nAμM−1; Fig. 4a). This could be dueto the reduction of DAQ, before it reaches the collector, by theAA that has not been completely eliminated by the largegenerator electrode configuration. This interaction would re-sult in an increased concentration gradient of DA at thegenerator, thus amplifying the anodic current. To be consistentwith this hypothesis, a decrease would be expected in theslope of the calibration curve for DA obtained at the collectorin the presence of 100 μM AA compared with DA alone.However, unlike the generator sensitivities with redox cyclingin the presence and absence of AA, the corresponding collec-tor sensitivities are not significantly different from each otherat the 95 % or 90 % confidence level, and thus that trendcannot be confirmed (0.192±0.011 nAμM−1 compared with0.209±0.034 nAμM−1, respectively). A decrease in the cali-bration slope at the generator electrode without redox cyclingwith AA compared with DA alone could not be confirmedeither, for the same reason (0.521±0.028 nAμM−1 vs 0.525±0.016 nAμM−1, respectively).

The y-intercept and the slope of the calibration curves inFig. 4b allow one to separate the contribution to anodiccurrent at the generator from oxidation of AA and DA,respectively. Thus, it is possible to obtain the collectionefficiency for DA in the presence of AA simply by takingthe ratio of the slope of the calibration curve for the mixtureat the collector to that for the mixture at the generator withredox cycling. The value (27.5±1.9 %) is significantly low-er (at the 90 % confidence level) than that in the absence ofAA (32.4±5.5 %). This is consistent with the hypothesisthat reduction of some DAQ by residual AA occurs before it

3864 A. Aggarwal et al.

Page 7: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

reaches the collector. Also supporting this hypothesis is theamplification factor for DA in the presence of AA, obtainedby taking the ratio of the slope of the calibration curve forthe mixture at the generator with redox cycling to that forthe mixture at the generator without redox cycling (1.34±0.08), which is significantly higher (at the 90 % confidencelevel) than that in the absence of AA (1.23±0.07). Asexpected, the collection efficiencies and amplification fac-tors for this electrode configuration for DA are still lowerthan those for an electrolyte solution containing only thereversible model compound Ru(NH3)6

3− (47.2 % and 1.46,respectively), whose fate is not affected by intramolecularrearrangements or homogeneous chemical reactions on thetimescale of the experiment.

The detection limits for DA in the presence and absenceof 100 μM AA at the electrodes with and without redox

cycling are listed in Table 1. In the presence of 100 μM AA,the current at the collector that is unique to DA provided asubmicromolar detection limit (0.454±0.026 μM). That val-ue is within the error of the detection limit for DA at thecollector in the absence of AA (0.417±0.067 μM), high-lighting the discriminating capability of redox cycling. Atthe generator, for comparison, the detection limit for DA inthe absence of AA with redox cycling (0.163±0.008 μM)was better than at the collector, as would be expected,because of its enhanced sensitivity. However, without redoxcycling, the detection limit at the generator for DA in theabsence of AA (0.107±0.003 μM) was even better, despitethe unamplified current and therefore lower sensitivity there.This unexpected result can be explained by the greater noiseat the generator when the collector is activated, and isreflected in a larger standard deviation of the blank signal,

Fig. 3 Examples of chronoamperometric responses are shown forredox cycling at the collector (a, d, g) and the generator (b, e, h) andwithout redox cycling at the generator (c, f, i) for solutions of artificialcerebrospinal fluid (aCSF) buffer (pH7.4) containing 10 μMDA (solidcurve), 25 μM DA (thick dashed curve), 50 μM DA (thin dashedcurve), and 100 μM DA (thick dotted curve) (a, b, c), 100 μM AA (d,e, f), and their mixtures (g, h, i) using the electrode configuration in

Fig. 1c. The generator was stepped to +0.500 V (from an equilibrationpotential of 0.000 V). For redox cycling, the collector’s potential wasstepped to −0.200 V (from an equilibration potential of −0.200 V). Inthe absence of redox cycling, the collector was left at open circuit. Allpotentials are referenced to Ag/AgCl (saturated KCl). (Note that thespikes in plot a at approximately 2 s and in plot d at approximately 4 sand approximately 6 s are from electronic noise.)

Detection of dopamine in the presence of excess ascorbic acid 3865

Page 8: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

which is used in the calculation of the limit of detection.(The standard deviations of the blank were 0.029 nA at thecollector, 0.035 nA at the generator with redox cycling, and0.019 nA at the generator without redox cycling.)

Although the submicromolar DA detection limits arequite respectable, compared with detection limits for redoxcycling previously reported in the literature, the detection

limits are not as optimal as they could be. One reason is alimitation of the potentiostat that was used here. Thepicoamp booster for the CHI 760 potentiostat does notsupport the secondary electrodes, and therefore does notfilter out noise at the collector. With instrumentation thatcan filter out noise at the electrodes, the detection limitsshould improve further. Also the collectors are held at apotential in the oxygen reduction regime. Thus, its reductionat the collector adds to error owing to variation fromexperiment to experiment, leading to poorer detectionlimits. One way to avoid this problem would be tolessen the overpotential at the collector electrode byapplying a less negative potential, e.g., 0.000 V versusAg/AgCl (saturated KCl).

Conclusions

Redox cycling with the unique MEA configuration de-scribed here was shown to be an effective electrochemicalmethod to detect DA and discriminate it from AA atphysiologically relevant concentrations in aCSF, withoutrequiring enhanced sensitivity through electrode modifica-tion, which can significantly decrease diffusion coeffi-cients and slow response times. The detection limitsreported here are sufficient to measure the submicromolarto micromolar concentrations of DA in vivo that havebeen found with other electrochemical detection methods[64]. However, redox cycling eliminates the need to sub-tract charging current, which restricts use of fast-scan CVto measuring DA transients. Thus, redox cycling is wellpoised for in vivo measurement of basal levels of DA inthe future.

There are several other electrochemically active com-pounds that can oxidize in a similar potential windowas DA and have various re-reduction behaviors at cath-odes. These include dihydroxyphenylacetic acid, seroto-nin, norepinephrine, epinephrine, and uric acid. In vitrostudies to detect DA in the presence of these and othercompounds using redox cycling are currently under way.

Ultimately, additional studies are needed to confirm thesuitability of redox cycling for in vivo DA measurements.Several interferences in natural samples that are difficult tomimic in vitro could affect the signal at the collector, in-cluding possible restricted diffusion in tissue [75], otherhomogeneous reactions that deplete DAQ, and current fromelectron transfer of other species at the cathodic collector.However, it is encouraging that DAQ can be detected withhigh concentrations of AA in vitro, suggesting thatadditional interfering reductants that are present in vivowould also need to have such high concentration orexhibit faster reaction kinetics to make a significantimpact on the collector signal from DAQ.

Fig. 4 Calibration curves of the steady-state chronoamperometricresponses as a function of DA concentration in solutions containing aaCSF buffer (pH7.4) and b 100 μM AA in aCSF buffer, shown for thegenerator current with (open squares, iG) and without (closed squares,iG/C) redox cycling, and for the collector current (closed circles, iC).The electrode configuration in Fig. 1c was used. The large y-intercept(35 nA) for the generator current in b, but not observed for thecollectors, is due to the oxidation of the 100 μMAA there. The averagecurrent taken at approximately 15 s from triplicate chronoamperomet-ric responses was used to produce the calibration curves. Error bars ateach marker represent ± one standard deviation

Table 1 The limit of detection (LOD) for dopamine (DA) in thepresence and absence of 100 μM ascorbic acid (AA) at the generatorand at the collector with and without redox cycling in artificial cere-brospinal fluid buffer, using chronoamperometry with the electrodeconfiguration in Fig. 1c. Each LOD was calculated as three times thestandard deviation of the blank signal, divided by the slope of thecalibration curve. The standard deviation of the LOD was obtainedfrom propagation of error

Generator withoutredox cycling

Generator withredox cycling

Collector withredox cycling

LOD (DA) 0.107±0.003 μM 0.163±0.008 μM 0.417±0.067 μM

LOD (DAwith AA)

0.108±0.006 μM 0.151±0.005 μM 0.454±0.026 μM

3866 A. Aggarwal et al.

Page 9: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

Investigations are also under way to scale down theMEA configuration onto probes having dimensions suit-able for insertion into brain tissue to implement redoxcycling for in vivo studies. The outer dimensions of theentire MEA used here were 140 μm in width and 2 mmin length. Simply compressing the overall width dimen-sions to 70 μm, by halving the gap and electrode widths,should not change the magnitude of the collector currentbecause the enhanced signal from the narrower gapmakes up for the loss in electrode area. (This assumesthat our electrode configuration would behave the sameway upon compression as an alternating configuration ofequally sized collector and generator electrodes [76].) Anelectroactive region that has a 70-μm width is similar tothat of previously fabricated silicon probes [59] and iswithin microfabrication capabilities. Decreasing the elec-trode length to 200 μm, a length similar to that of CFEsused for in vivo studies, would decrease the cathodiccurrent at the collector by a factor of 10 (to approxi-mately 0.1 nA for 1 μM DA in 100 μm AA). Improve-ments in instrumentation with filtering and propershielding to decrease electronic noise would allow similaror better detection limits to be obtained. This should beachievable, given that instrumentation used for fast-scanCV in vivo already allows magnitudes of faradaic currentof 0.1–10 nA to be successfully measured.

MEA devices with a smaller gap between generators andcollectors can be fabricated to offer further improvements.First, the time for DAQ and DA to diffuse across the gapshould decrease dramatically, because it depends on thesquare of the distance (Eq. 1). One benefit is that the timefor the collector to respond to changes in DA concentrationwould also decrease dramatically. Another is that longerresponse times from slowed mass transport in polymers thatcould be used to modify a MEA may be offset by the shorterdistances and yield a response time similar to that of a bareMEA. Thus, coatings (thinner than the gap) may be consid-ered, especially for long-term use in vivo, to eliminateadditional interferences and protect electrodes from fouling.Second, a narrower gap would cause an increase in themagnitude of the collector signal, not only because of anincreased concentration gradient of DAQ across the gap, buteven more so because less DAQ would be lost throughintramolecular cyclization (k=0.13±0.05 s−1 at physiologi-cal pH) [77, 78] and reactions with reductants (such asresidual AA) in that space.

Acknowledgments Funding was provided in part by the NationalScience Foundation (CHE-0719097) and the Arkansas BiosciencesInstitute, the major research component of the Arkansas TobaccoSettlement Proceeds Act of 2000. We thank Errol Porter for adviceon microfabrication. The use of the High Density Electronics Centermicrofabrication facilities is also acknowledged. We express ourappreciation to Adrian Michael of the University of Pittsburgh for

insightful discussions about applications. T.S. Hollingsworth isacknowledged for assistance in preparing this manuscript.

References

1. Benes FM (2001) Arvid Carlsson and the discovery of dopamine.Trends Pharmacol Sci 22(1):46–47

2. Berridge KC, Robinson TE (1998) What is the role of dopamine inreward: hedonic impact, reward learning, or incentive salience?Brain Res Rev 28(3):309–369

3. Pan WX, Schmidt R, Wickens JR, Hyland BI (2005) Dopaminecells respond to predicted events during classical conditioning:evidence for eligibility traces in the reward-learning network. JNeurosci 25(26):6235–6242

4. Tobler PN, Fiorillo CD, Schultz W (2005) Adaptive coding ofreward value by dopamine neurons. Science 307(5715):1642–1645

5. Wise RA (1998) Drug-activation of brain reward pathways. DrugAlcohol Depend 51(1–2):13–22

6. Redgrave P, Gurney K (2006) The short-latency dopamine signal:a role in discovering novel actions? Nat Rev Neurosci 7(12):967–975

7. Salgado-Pineda P, Delaveau P, Blin O, Nieoullon A (2005)Dopaminergic contribution to the regulation of emotional perception.Clin Neuropharmacol 28(5):228–237

8. Schultz W (2007) Multiple dopamine functions at different timecourses. Annu Rev Neurosci 30:259–288

9. Obata T (2002) Dopamine efflux by MPTP and hydroxyl radicalgeneration. J Neural Transm 109(9):1159–1180

10. Wightman RM, May LJ, Michael AC (1988) Detection of dopaminedynamics in the brain. Anal Chem 60(13):769A–793A

11. Bibb JA, Yan Z, Svenningsson P, Snyder GL, Pieribone VA,Horiuchi A, Nairn AC, Messer A, Greengard P (2000) Severedeficiencies in dopamine signaling in presymptomatic Hunting-ton’s disease mice. Proc Natl Acad Sci U S A 97(12):6809–6814

12. Phillips PEM, Stuber GD, Heien M, Wightman RM, Carelli RM(2003) Subsecond dopamine release promotes cocaine seeking.Nature 422(6932):614–618

13. Koob GF, Bloom FE (1988) Cellular and molecular mechanisms ofdrug-dependence. Science 242(4879):715–723

14. Mink JW (2001) Basal ganglia dysfunction in Tourette’ssyndrome: a new hypothesis. Pediatr Neurol 25(3):190–198

15. Grace AA (1991) Phasic versus tonic dopamine release and themodulation of dopamine system responsivity–a hypothesis for theetiology of schizophrenia. Neuroscience 41(1):1–24

16. Salahpour A, Ramsey AJ, Medvedev IO, Kile B, Sotnikova TD,Holmstrand E, Ghisi V, Nicholls PJ, Wong L, Murphy K, SesackSR, Wightman RM, Gainetdinov RR, Caron MG (2008) Increasedamphetamine-induced hyperactivity and reward in mice overex-pressing the dopamine transporter. Proc Natl Acad Sci U S A105(11):4405–4410

17. Garris PA, Rebec GV (2002) Modeling fast dopamine neuro-transmission in the nucleus accumbens during behavior. BehavBrain Res 137:47–63

18. Sunsay C, Rebec GV (2008) Real-time dopamine efflux in thenucleus accumbens core during pavlovian conditioning. BehavNeurosci 122(2):358–367

19. Aragona BJ, Cleaveland NA, Stuber GD, Day JJ, Carelli RM,Wightman RM (2008) Preferential enhancement of dopaminetransmission within the nucleus accumbens shell by cocaine isattributable to a direct increase in phasic dopamine release events.J Neurosci 28(35):8821–8831

20. Zhang T, Zhang L, Liang Y, Siapas AG, Zhou F-M, Dani JA(2009) Dopamine signaling differences in the nucleus accumbens

Detection of dopamine in the presence of excess ascorbic acid 3867

Page 10: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

and dorsal striatum exploited by nicotine. J Neurosci 29(13):4035–4043

21. Borland LM, Michael AC (2004) Voltammetric study of thecontrol of striatal dopamine release by glutamate. J Neurochem91(1):220–229

22. Oneill RD (1994) Microvoltammetric techniques and sensors formonitoring neurochemical dynamics in-vivo–a review. Analyst119(5):767–779

23. Robinson DL, Hermans A, Seipel AT, Wightman RM (2008)Monitoring rapid chemical communication in the brain. ChemRev 108(7):2554–2584

24. Borland LM, Shi GY, Yang H, Michael AC (2005) Voltammetricstudy of extracellular dopamine near microdialysis probes acutelyimplanted in the striatum of the anesthetized rat. J Neurosci Methods146(2):149–158

25. Kulagina NV, Zigmond MJ, Michael AC (2001) Glutamateregulates the spontaneous and evoked release of dopamine inthe rat striatum. Neuroscience 102(1):121–128

26. Bungay PM, Newton-Vinson P, Isele W, Garris PA, Justice JB(2003) Microdialysis of dopamine interpreted with quantitativemodel incorporating probe implantation trauma. J Neurochem86(4):932–946

27. Gonon F, Buda M, Cespuglio R, Jouvet M, Pujol JF (1980) In vivoelectrochemical detection of catechols in the neostriatum of anesthetizedrats–dopamine or dopac. Nature 286(5776):902–904

28. Gonon F, Buda M, Cespuglio R, Jouvet M, Pujol JF (1981)Voltammetry in the striatum of chronic freely moving rats: detectionof catechols and ascorbic acid. Brain Res 223(1):69–80

29. Popa E, Notsu H, Miwa T, Tryk DA, Fujishima A (1999) Selectiveelectrochemical detection of dopamine in the presence of ascorbicacid at anodized diamond thin film electrodes. Electrochem SolidState Lett 2(1):49–51

30. Rice ME (2000) Ascorbate regulation and its neuroprotective rolein the brain. Trends Neurosci 23(5):209–216

31. Adams RN, Conti J, Marsden CA, Strope E (1978) Measurementof dopamine and 5-hydroxytryptamine release in CNS of freelymoving unanesthetized rats. Br J Pharmacol 64(3):P470–P471

32. Hoffman AF, Gerhardt GA (1998) In vivo electrochemical studiesof dopamine clearance in the rat substantia nigra: effects of locallyapplied uptake inhibitors and unilateral 6-hydroxydopaminelesions. J Neurochem 70(1):179–189

33. Miller AD, Forster GL, Yeomans JS, Blaha CD (2005) Midbrainmuscarinic receptors modulate morphine-induced accumbal andstriatal dopamine efflux in the rat. Neuroscience 136(2):531–538

34. Unger EL, Eve DJ, Perez XA, Reichenbach DK, Xu YQ, Lee MK,Andrews AM (2006) Locomotor hyperactivity and alterations indopamine neurotransmission are associated with overexpression ofA53T mutant human alpha-synuclein in mice. Neurobiol Dis21(2):431–443

35. Venton BJ, Troyer KP, Wightman RM (2002) Response times ofcarbon fiber microelectrodes to dynamic changes in catecholamineconcentration. Anal Chem 74(3):539–546

36. Batchelor-McAuley C, Dickinson EJF, Rees NV, Toghill KE,Compton RG (2012) New electrochemical methods. Anal Chem84(2):669–684

37. Millar J, Stamford JΑ, Kruk ZL, Wightman RM (1985) Electrochemi-cal, pharmacological and electrophysiological evidence of rapid dopa-mine release and removal in the rat caudate nucleus following electricalstimulation of the median forebrain bundle. Eur J Pharm 109:341–348

38. Wipf DO, Kristensen EW, Deakin MR, Wightman RM (1988)Fast-scan cyclic voltammetry as a method to measure rapid het-erogeneous electron-transfer kinetics. Anal Chem 60(4):306–310

39. Perone SP, Kretlow WJ (1966) Application of controlled potentialtechniques to study of rapid succeeding chemical reaction coupledto electro-oxidation of ascorbic acid. Anal Chem 38(12):1760–1763

40. Akkermans RP, Wu M, Bain CD, Fidel-Suarez M, Compton RG(1998) Electroanalysis of ascorbic acid: a comparative study oflaser ablation voltammetry and sonovoltammetry. Electroanalysis10(9):613–620

41. Borsook H, Davenport HW, Jeffreys CEP, Warner RC (1937) Theoxidation of ascorbic acid and its reduction in vivo and in vitro. JBiol Chem 117(1):237–279

42. Niwa O, Morita M, Tabei H (1991) Highly sensitive and selectivevoltammetric detection of dopamine with vertically separated in-terdigitated array electrodes. Electroanalysis 3(3):163–168

43. Peng WF, Wang EK (1993) Preparation and characterization of amulticylinder microelectrode coupled with a conventional glassy-carbon electrode and Its application to the detection of dopamine.Anal Chim Acta 281(3):663–671

44. Cullison JK, Waraska J, Buttaro DJ, Acworth IN, Bowers ML(1999) Electrochemical detection of catecholamines at sub-5 fglevels by redox cycling. J Pharm Biomed Anal 19(1–2):253–259

45. Dam VAT, Olthuis W, van den Berg A (2007) Redox cycling withfacing interdigitated array electrodes as a method for selectivedetection of redox species. Analyst 132(4):365–370

46. Niwa O, Kurita R, Liu ZM, Horiuchi T, Torimitsue K (2000)Subnanoliter volume wall-jet cells combined with interdigitatedmicroarray electrode and enzyme modified planar microelectrode.Anal Chem 72(5):949–955

47. Niwa O, Morita M (1996) Carbon film-based interdigitated ring arrayelectrodes as detectors in radial flow cells. Anal Chem 68(2):355–359

48. Niwa O, Morita M, Tabei H (1994) Highly selective electrochemicaldetection of dopamine using interdigitated array electrodes modifiedwith Nafion polyester ionomer layered film. Electroanalysis6(3):237–243

49. Vandaveer WR IV, Woodward DJ, Fritsch I (2003) Redox cyclingmeasurements of a model compound and dopamine in ultrasmallvolumes with a self-contained microcavity device. ElectrochimActa 48(20–22):3341–3348

50. Niwa O, Tabei H (1994) Voltammetric measurements of reversibleand quasi-reversible redox species using carbon-film basedinterdigitated array microelectrodes. Anal Chem 66(2):285–289

51. Katelhon E, Hofmann B, Lemay SG, ZevenbergenMAG,OffenhausserA, Wolfrum B (2010) Nanocavity redox cycling sensors forthe detection of dopamine fluctuations in microfluidic gradients.Anal Chem 82(20):8502–8509

52. Liu ZM, Niwa O, Kurita R, Horiuchi T (2000) Carbon film-basedinterdigitated array microelectrode used in capillary electrophoresiswith electrochemical detection. Anal Chem 72(6):1315–1321

53. Zachek MK, Takmakov P, Park J, Wightman RM, McCarty GS(2010) Simultaneous monitoring of dopamine concentration atspatially different brain locations in vivo. Biosens Bioelectron25:1179–1185

54. Huang XJ, O'Mahony AM, Compton RG (2009) Microelectrodearrays for electrochemistry: approaches to fabrication. Small5(7):776–788

55. Bruno JP, Gash C, Martin B, Zmarowski A, Pomerleau F,Burmeister J, Huettl P, Gerhardt GA (2006) Second-by-secondmeasurement of acetylcholine release in prefrontal cortex. EurJ Neurosci 24(10):2749–2757

56. Bruno JP, Sarter M, Gash C, Parikh V (2006) Choline- andacetylcholine-sensitive microelectrodes. Encyclopedia of sensors,vol 2. American Scientific Publishers, Valencia

57. Burmeister JJ, Pomerleau F, Huettl P, Gash CR, Wemer CE, BrunoJP, Gerhardt GA (2008) Ceramic-based multisite microelectrodearrays for simultaneous measures of choline and acetylcholine inCNS. Biosens Bioelectron 23(9):1382–1389

58. Zachek MK, Park J, Takmakov P, Wightman RM, McCarty GS(2010) Microfabricated FSCV-compatible microelectrode array forreal-time monitoring of heterogeneous dopamine release. Analyst135:1556–1563

3868 A. Aggarwal et al.

Page 11: Detection of dopamine in the presence of excess ascorbic acidat physiological concentrations through redox cyclingat an unmodified microelectrode array

59. Sreenivas G, Ang S, Fritsch-Faules I, Brown WD, GerhardtGA, Woodward DJ (1996) Fabrication and characterization ofsputtered-carbon microelectrode arrays. Anal Chem 68:1858–1864

60. Peters JL, Miner LH, Michael AC, Sesack SR (2004) Ultrastructureat carbon fiber microelectrode implantation sites after acutevoltammetric measurements in the striatum of anesthetizedrats. J Neurosci Methods 137(1):9–23

61. Clapp-Lilly KL, Roberts RC, Duffy LK, Irons KP, Hu Y, Drew KL(1999) An ultrastructural analysis of tissue surrounding amicrodialysis probe. J Neurosci Methods 90(2):129–142

62. Mitala CM, Wang YX, Borland LM, Jung M, Shand S, Watkins S,Weber SG, Michael AC (2008) Impact of microdialysis probeson vasculature and dopamine in the rat striatum: a combinedfluorescence and voltammetric study. J Neurosci Methods174(2):177–185

63. Zhou F, Zhu XW, Castellani RJ, Stimmelmayr R, Perry G, SmithMA, Drew KL (2001) Hibernation, a model of neuroprotection.Am J Pathol 158(6):2145–2151

64. Michael AC, Borland LM, Mitala JJ, Willoughby BM, MotzkoCM (2005) Theory for the impact of basal turnover on dopa-mine clearance kinetics in the rat striatum after medial fore-brain bundle stimulation and pressure ejection. J Neurochem94(5):1202–1211

65. Burmeister JJ, Moxon K, Gerhardt GA (2000) Ceramic-basedmultisite microelectrodes for electrochemical recordings. AnalChem 72:187–192

66. Burmeister JJ, Pomerleau F, Palmer M, Day BK, Huettl P, GerhardtGA (2002) Improved ceramic-based multisite microelectrode forrapid measurements of L-glutamate in the CNS. J NeurosciMethods 119:163–171

67. Kume-Kick J, Rice ME (1998) Dependence of dopaminecalibration factors on media Ca2+ and M2+ at carbon-fibermicroelectrodes used with fast-scan cyclic voltammetry. JNeurosci Methods 84:55–62

68. Rice M, Nicholson C (1989) Measurement of nanomolar dopaminediffusion using low-noise perfluorinated ionomer coated carbon

fiber microelectrodes and high-speed cyclic voltammetry. AnalChem 61:1805–1810

69. Wiedemann D, Kawagoe K, Kennedy R, Ciolkowski E, WightmanR (1991) Strategies for low detection limit measurements withcyclic voltammetry. Anal Chem 63:2965–2970

70. Zachek MK, Hermans A, Wightman RM, McCarty GS (2008)Electrochemical dopamine detection: comparing gold and carbonfiber microelectrodes using background subtracted fast scan cyclicvoltammetry. J Electroanal Chem 614(1):113–120

71. Aggarwal A (2011) Studies toward the development of amicroelectrode array for detection of dopamine through redoxcycling. PhD thesis, University of Arkansas, Fayetteville

72. Bard AJ, Crayston JA, Kittlesen GP, Shea TV, Wrighton MS(1986) Digital simulation of the measured electrochemicalresponse of reversible redox couples at microelectrode arrays:consequences arising from closely spaced ultramicroelectrodes.Anal Chem 58(11):2321–2331

73. Bard AJ, Faulkner LR (1980) Electrochemical methods: fundamentalsand applications. Wiley, New York

74. Lu Y, Peters JL, Michael AC (1998) Direct comparison of theresponse of voltammetry and microdialysis to electrically evokedrelease of striatal dopamine. J Neurochem 70(2):584–593

75. Amatore C, Kelly RS, Kristensen EW, Kuhr WG, Wightman RM(1986) Effects of restricted diffusion at ultramicroelectrodes inbrain tissue: the pool model: theory and experiment for chronoam-perometry. J Electroanal Chem 213(1):31–42

76. Aoki K, Morita M, Niwa, Tabei H (1988) Quantitative analysis ofreversible diffusion-controlled currents of redox soluble species atinterdigitated array electrodes under steady-state conditions. JElectroanal Chem 256:269–282

77. Hawley MD, Tatawawa S, Piekarsk S, Adams RN (1967) Electro-chemical studies of oxidation pathways of catecholamines. J AmChem Soc 89(2):447–450

78. Ciolkowski EL, Cooper BR, Jankowski JA, Jorgenson JW,WightmanRM (1992) Direct observation of epinephrine and norepinephrinecosecretion from individual adrenal-medullary chromaffin cells. JAm Chem Soc 114(8):2815–2821

Detection of dopamine in the presence of excess ascorbic acid 3869