throbber
Water Research 36 (2002) 85–94
`
`Electrochemical generation of hydrogen peroxide from
`dissolved oxygen in acidic solutions
`
`Zhimin Qiang, Jih-Hsing Chang, Chin-Pao Huang*
`
`Department of Civil and Environmental Engineering, University of Delaware, 137, DuPont Hall, Newark, DE 19716-3120, USA
`
`Received 1 July 2000; received in revised form 1 April 2001; accepted 1 April 2001
`
`Abstract
`
`Hydrogen peroxide (H2O2) was electro-generated in a parallel-plate electrolyzer by reduction of dissolved oxygen
`(DO) in acidic solutions containing dilute supporting electrolyte. Operational parameters such as cathodic potential,
`oxygen purity and mass flow rate, cathode surface area, pH,
`temperature, and inert supporting electrolyte
`concentration were systematically investigated as to improve the Faradic current efficiency of H2O2 generation.
`Results indicate that significant self-decomposition of H2O2 only occurs at high pH (>9) and elevated temperatures
`(>231C). Results also indicate that the optimal conditions for H2O2 generation are cathodic potential of –0.5 V vs.
`saturated calomel electrode (SCE), oxygen mass flow rate of 8.2  102 mol/min, and pH 2. Under the optimal
`conditions, the average current density and average current efficiency are 6.4 A/m2 and 81%, respectively. However,
`when air is applied at the optimal flow rate of oxygen, the average current density markedly decreases to 2.1 A/m2, while
`the average current efficiency slightly increases to 90%. The limiting current density is 6.4 A/m2, which is independent of
`cathode geometry and surface area. H2O2 generation is favored at low temperatures. In the concentration range studied
`(0.01–0.25 M), the inert supporting electrolyte (NaClO4) affects the total potential drop of the electrolyzer, but does not
`affect the net generation rate of H2O2. r 2001 Elsevier Science Ltd. All rights reserved.
`
`Keywords: Hydrogen peroxide; Electrochemical generation; Oxygen; Air; Acidic solutions
`
`1. Introduction
`
`is an environmentally
`Hydrogen peroxide (H2O2)
`leaves no hazardous
`friendly chemical because it
`residuals but oxygen and water after reaction. It has
`been widely applied to the synthesis of organic
`compounds, bleaching of paper pulp,
`treatment of
`wastewater, and destruction of hazardous organic
`wastes. In the environmental field, H2O2 is used as a
`supplement of oxygen source to enhance the bioreme-
`diation of contaminated aquifers [1,2]. Moreover, H2O2
`coupled with ozone or UV radiation can effectively
`decompose aqueous organic contaminants [3–6]. The
`most common environmental application of H2O2 is the
`
`*Corresponding author. Tel.: +1-302-831-2442; fax: +1-
`302-831-3640.
`E-mail address: huang@ce.udel.edu (Chin-Pao Huang).
`
`Fenton’s reagent, an aqueous mixture of H2O2 and
`Fe2+. Under an acidic condition, the reaction between
`H2O2 and Fe2+ generates hydroxyl radicals that are
`strong enough to non-selectively oxidize most organic as
`well as some inorganic compounds. The Fenton’s
`reagent has recently been applied to in-situ remediation
`of contaminated soils and groundwaters [7–9].
`H2O2 is usually produced by electrochemical methods,
`such as electrolysis of inorganic chemicals (H2S2O8,
`KHSO4 and NH4HSO4) and autoxidation of organic
`compounds (alkylhydroanthraquinones and isopropyl
`alcohol) [10]. The electrolysis of inorganics requires
`excessive energy and the autoxidation of organics
`requires non-aqueous solvents for catalyst cycle [11].
`H2O2 may also be directly generated from water,
`hydrogen, and oxygen using thermal, photochemical
`and electrical discharge processes. However,
`these
`processes require specific operational conditions such
`
`0043-1354/01/$ - see front matter r 2001 Elsevier Science Ltd. All rights reserved.
`PII: S 0 0 4 3 - 1 3 5 4 ( 0 1 ) 0 0 2 3 5 - 4
`
`Tennant Company
`Exhibit 1120
`
`

`

`86
`
`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`as high temperature, combustion, UV radiation plus
`mercury vapor, or high voltage. In recent years, small
`scale, on-site H2O2 production processes have gained
`increasing attention because of the cost and the hazards
`associated with the transport and handling of commer-
`cial concentrated H2O2 [11]. If H2O2 can be generated
`on-site in an economic and safe way, its field application
`will be largely simplified. For example,
`it would be
`attractive to use the on-site generated H2O2, coupled
`with ozone, UV, or Fe2+, for the detoxification of
`effluents from electro-kinetics, pump and treat, soil
`washing, and soil flushing processes.
`Most electro-generation of H2O2 experiments are
`conducted in alkaline solutions with a high electrolyte
`concentration for the purpose of bleaching paper pulp
` is
`[12–19]. In concentrated alkaline solutions, HO2
`formed (pKa; H2O2 ¼ 11:62 at 251C) which will be
`immediately repelled by the cathode upon its generation.
` to OH is minimized, a
`Because the reduction of HO2
`high current efficiency (about 80–95%) can usually be
`achieved. However,
`if H2O2 is generated in alkaline
`solutions, a substantial amount of acid will be consumed
`for pH adjustment if an acidic condition is required. The
`Fenton’s reagent, which is most commonly applied in
`organic synthesis and effluent treatment, has an optimal
`pH range of 2.5–3.5. Moreover, a high electrolyte
`concentration not only increases the treatment cost,
`but also introduces additional pollutants. Therefore, it is
`desirable to generate H2O2 in acidic solutions only
`containing dilute supporting electrolyte.
`It has been reported that H2O2 can be electrochemi-
`cally generated by reduction of dissolved oxygen (DO)
`in acidic solutions. The H2O2 so generated can be
`coupled with Fe2+ to produce the Fenton’s reagent for
`either degradation or synthesis of organic compounds
`[20–28]. In order to differentiate this process from
`conventional Fenton process that uses commercial
`H2O2, the term ‘‘electro-Fenton process’’
`is applied.
`The major advantages of the electro-Fenton process
`include: (1) H2O2 can be continuously generated on-site
`whenever needed, which eliminates acquisition, ship-
`ment and storage; (2) A dilute H2O2 solution enhances
`safety during material handling; (3) The production
`process can be simply conducted at ambient pressure
`and temperature; (4) Fe2+ can be electrochemically
`regenerated at
`the cathode, which minimizes
`the
`quantity of iron sludge; and (5) Oxygen or air sparging
`enhances the mixing of reaction solution. The disadvan-
`tage is that H2O2 will accumulate at the cathode-
`solution interface and may be partially decomposed.
`Protons at a high concentration may also compete for
`electrons,
`leading to hydrogen gas evolution. Both
`effects will
`reduce the current efficiency of H2O2
`production. Therefore,
`in acidic solutions, cathodic
`potential and solution pH are two essential factors
`controlling the current efficiency.
`
`Though the electro-generation of H2O2 in acidic
`solutions has been studied by a few researchers, there
`exist conflicting results. Sudoh et al. [21] investigated the
`decomposition of aqueous phenol by electro-generated
`Fenton’s reagent. They found that the highest current
`efficiency (85%) was obtained at a cathodic potential of
`0.6 V vs. a saturated Ag/AgCl electrode (SSE) and pH
`3. Tzedakis et al. [23] reported that the electrolysis of an
`oxygen-saturated H2SO4 solution (0.6 M) using a stirred
`mercury pool electrode yielded a current efficiency of
`55% at a cathodic potential of 0.30 V vs. SCE. Chu
`[27] used the electro-Fenton process to remove aqueous
`chlorophenols at a cathodic potential of 0.6 V vs. SCE,
`and reported that the current efficiency increased with
`decreasing pH. Hsiao and Nobe [25] investigated the
`oxidative hydroxylation of phenol and chlorobenzene
`using electro-generated Fenton’s reagent. They reported
`that the optimal cathodic potential was 0.55 V vs.
`SCE, and the generation of H2O2 was favored at low pH
`values.
`The primary objective of the present study is to
`improve the Faradic current efficiency of H2O2 genera-
`tion in acidic solutions. It is also expected that the
`results will clarify existing discrepancies among various
`studies of
`its kind. Influential parameters such as
`cathodic potential, oxygen purity and mass flow rates,
`cathode surface area, solution pH, temperature, and
`inert supporting electrolyte concentration were system-
`atically examined. Considering that a high electrolyte
`concentration is usually not feasible for effluent treat-
`ment by the electro-generated H2O2, the inert support-
`ing electrolyte, i.e., NaClO4, was used only at a low
`concentration.
`
`2. Experimental
`
`Fig. 1 shows the schematic diagram of the reaction
`system. An acrylic parallel-plate electrolyzer was em-
`ployed for the electro-generation of H2O2. The cathodic
`and anodic compartments had a volume of 4.50 and
`3.15 l, respectively. During the experiments, 4.0 l of
`catholyte and 3.0 l of anolyte were used and both were
`completely mixed by a magnetic stir plate. A cation
`exchange membrane (Neosepta CMX, Electrosynthesis
`Company, Lancaster, NY) was used to separate the two
`compartments. This membrane prohibits the penetra-
`tion of anions and H2O2 molecules, but allows cations to
`freely penetrate through it. As a result, H2O2 generated
`at the cathode will be confined in the catholyte, avoiding
`its decomposition at the anode. Moreover, protons
`generated at the anode will be electrically driven to the
`catholyte, partially supplementing the protons con-
`sumed for H2O2 synthesis. Sodium perchlorate (Na-
`ClO4) was used as an inert supporting electrolyte (or
`background ionic strength). Both cathode and anode
`
`

`

`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`87
`
`Fig. 1. Schematic diagram of reaction system.
`
`were made of corrosion-resistant graphite (Grade 2020,
`Carbon of America, Bay City, MI). Sudoh et al. [16]
`reported that graphite was the best cathode material for
`the electro-generation of H2O2 in alkaline solutions,
`while metal cathodes such as copper, stainless steel, lead
`and nickel were likely to decompose H2O2. Three
`cathode
`geometries
`(Fig. 1),
`i.e.,
`plain
`plate
`(17.78 cm  15.24 cm, or 7 in  6 in), plate with pro-
`truded short ‘‘fingers’’ (17.78 cm  15.24 cm  1.02 cm,
`or 7 in  6 in  0.4 in) and plate with protruded long
`(17.78 cm  15.24 cm  1.52 cm,
`‘‘fingers’’
`or
`7 in
` 6 in  0.6 in) were employed to investigate the effect
`of cathode surface area. Unless otherwise stated, all
`experiments were conducted using the long-finger plate.
`A graphite plain plate (8.89 cm  15.24 cm, or 3.5 in
` 6 in) was used as the anode. Copper wires were
`connected to both electrodes through Teflon screws. The
`connections were carefully sealed with silicon to prevent
`copper
`electro-corrosion. Compressed oxygen gas
`(99.6%) and air were used as DO sources. The gas was
`sparged into the catholyte through a porous pipe-
`diffuser placed right under the cathode. The catholyte
`was pre-saturated with DO by purging pure oxygen gas
`or air for 15 min before electrolysis was initiated. The
`catholyte pH was controlled by a pH-stat (Model pH-
`40, New Brunswick Scientific Co., Edison, NJ), HClO4
`(1 M) and NaOH (1 M) solutions. A high performance
`combination pH probe (Cat. No. 376490, Corning Inc.,
`Corning, NY) was placed behind the cathode to avoid
`the interference from the electrical field. The solution
`
`temperature was controlled by a thermostat (Model EX-
`200, Brookfield Engineering Laboratories, Inc., Stought-
`on, MA) and a water bath.
`Polarization curves were obtained by cyclic voltam-
`metry using a three-electrode bi-potentiostat (Model
`AFRDE4, Pine Instrument Co., Grove City, PA). A
`saturated calomel electrode (SCE) was used as the
`reference electrode. The SCE was inserted into a Luggin
`capillary filled with saturated KCl solution. By placing
`the tip of the Luggin capillary in contact with the
`cathode surface, the cathodic potential can be accurately
`controlled against the SCE. The cathodic potential of
`the parallel-plate electrolyzer was swept from 0 to –0.8 V
`(vs. SCE) at a linear rate of 33.3 mV/s. Transient current
`response was recorded by a Hewlett Packard X-Y
`recorder (Model 7001A, Moseley Division, Pasadena,
`CA). The electro-generation of H2O2 experiments were
`carried out under either constant potential or constant
`current mode. Usually, constant potential mode is used
`to derive fundamental electrochemical information in
`laboratory scale experiments, while constant current
`mode is commonly used in industrial electrolysis because
`it is technically much easier to control the current than
`the potential [29].
`In this study, the factors significantly affecting the
`limiting current of H2O2 generation, including cathodic
`potential, oxygen purity and mass flow rate, and cathode
`surface area were investigated using the constant
`potential mode. Under this mode, the electrical current
`was monitored on-line by a digital multimeter (Model
`
`

`

`88
`
`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`22-183A, Tanday Co., Fort Worth, Texas). The effect of
`pH was studied using both constant potential and
`constant current modes. The effects of temperature and
`inert supporting electrolyte concentration were investi-
`gated by the constant current mode. A regulated DC
`power
`supply (Model WP-705B, Vector-Vid Inc.,
`Horsham, PA) was employed to provide constant
`current. The concentration of H2O2 was determined by
`the titanic sulfate [Ti(SO4)2] method [16]. A diode array
`spectrophotometer (Model 8452A, Hewlett Packard)
`was used to measure the light absorbance of the Ti4+–
`H2O2 orange complex at 410 nm. The DO concentration
`was determined by an oxygen electrode (Model 97-08-
`99, Orion Research Inc., Beverly, MA). A pH meter was
`used to record the DO concentration in the range of 0–
`14 mg/l. When the concentration exceeded the upper
`response limit, dilution was made with deoxygenated
`distilled water.
`
`3. Results and discussion
`
`3.1. Stability of hydrogen peroxide
`
`In a thoroughly clean container without the presence
`of any catalysts, H2O2 is very stable at any concentra-
`tion. However, the presence of a trace amount of metal
`ions in the solution or on the container surface will lead
`to the decomposition of H2O2. In practice, stabilizing
`agents such as sodium stannate, 8-hydroxyquinoline and
`sodium pyrophosphate are commonly used to stabilize
`H2O2 for long-term storage [10]. The self-decomposition
`rate of H2O2
`is primarily influenced by pH and
`temperature. Sudoh et al. [16] attributed the low current
`efficiency (6.88%) of H2O2 generation in alkaline
`solutions at 301C to a high self-decomposition rate of
`H2O2. Solution pH influences the chemical speciation of
`both H2O2 and trace metals. The trace metals that can
`catalytically initiate the self-decomposition of H2O2 may
`be introduced from acid and base used for pH
`adjustment, from container surfaces, and even from
`distilled water. At low pH, H2O2 and free metal ions
` and metal-hydroxo
`predominate. At high pH, HO2
`complexes are the major
`species. Therefore,
`it
`is
`necessary to investigate the effects of pH and tempera-
`ture on the self-decomposition of H2O2.
`The H2O2 solution with an initial concentration of
`150 mg/l was prepared by diluting a commercial grade
`H2O2 solution (31.5% by weight) with distilled water. A
`series of plastic bottles (250 ml) were washed with 1 M
`HClO4 solution, and then filled with the H2O2 solution.
`The pH was adjusted by reagent grade HClO4 or NaOH
`to cover the range of 1–13. The effect of temperature was
`investigated at 101C, 231C and 501C. At selected time
`intervals, the concentration of H2O2 was determined.
`Fig. 2 shows the self-decomposition of H2O2 at various
`
`Fig. 2. Stability of hydrogen peroxide.
`
`pH values, temperatures and reaction times. Results
`relatively stable at pHo9.
`indicate that H2O2
`is
`However, above pH 9, H2O2 decomposes markedly
`with increasing pH, temperature and reaction time.
`There is complete H2O2 decomposition at pH 13 and
`temperature 501C after 96 h. According to Schumb et al.
`[10], even with the purest H2O2 and at elevated
`temperatures, the decomposition of H2O2 in the liquid
`phase is not a homogeneous autodecomposition process
`increasing temperature
`of the H2O2 itself. Generally,
`increases the reaction rate. The self-decomposition of
`H2O2 at high pH and elevated temperatures are
`attributed in part to the catalytic effect of the container
`walls and the reagent impurities. Another aspect is the
` in the base catalyzedanion, HO2. The role of HO2
`
`
`H2O2 decomposition was
`suggested by Abel
`[30]
`following the reaction:
`H2O2 þ HO
`-H2O þ O2 þ OH:
`
`2
`However, a low temperature (e.g., 101C) suppresses
`H2O2 self-decomposition, even at high pH. It is also
`noted that
`in the acidic region,
`the highest
`self-
`decomposition rate appears at pH 3. It is known that
`pH 3 is the optimal value for the Fenton’s reagent. The
`possible presence of trace metals in the solution or on
`the bottle surfaces may catalytically stress the decom-
`position of H2O2 at pH 3. It is noted that the electrolysis
`was conducted in acidic conditions, i.e., pHp4.0, and
`the electrolysis time was 2 h. Therefore, based on the
`results shown in Fig. 2,
`it
`is clear that
`the self-
`decomposition of H2O2 would be insignificant.
`
`ð1Þ
`
`3.2. Optimal cathodic potential
`
`In acidic solutions, the dissolved oxygen is electro-
`chemically reduced to H2O2 at the cathode
`O2 þ 2Hþ þ 2e-H2O2; Eo ¼ 0:440 V vs: SCE:
`
`ð2Þ
`
`

`

`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`89
`
`continuously increases with increasing H2O2 concentra-
`tion.
`In the limiting current region, H2O2 generation is
`controlled by the mass transfer of DO through the
`cathode-solution diffusion layer, rather than by the
`electron transfer between DO and cathode. Since the
`DO concentration at
`the cathode surface rapidly
`approaches zero after electrolysis starts, the limiting
`current under a steady-state condition can be expressed
`by the following equation for macroscopic electrodes
`[29]:
`IL ¼ kmnFAeC *;
`where IL represents the limiting current ðAÞ; km is the
`mass transfer coefficient (m/s), n is the stoichiometric
`number of electrons transferred, F is the Farady’s
`constant (96,490 C/mol), Ae
`is the effective cathode
`surface area ðm2Þ; and C is the DO concentration in
`bulk solution ðMÞ: The mass transfer coefficient, km; can
`be determined by Eq. (7):
`km ¼ D=d
`
`ð6Þ
`
`ð7Þ
`
`Two side reactions simultaneously occur at the cathode:
`(1) the reduction of H2O2 to H2O due to the accumula-
`tion of H2O2 at the cathode-solution interface, and (2)
`the hydrogen gas evolution. These reactions are shown
`as follows:
`H2O2þ2Hþþ2e-2H2O; Eo ¼ 1:534 V vs: SCE;
`
`ð3Þ
`
`2Hþþ2e-H2; Eo ¼ 0:242 V vs: SCE:
`
`ð4Þ
`
`At the anode, the oxidation of H2O releases oxygen gas
`and protons
`2H2O-4HþþO2þ4e; Eo ¼ 0:987 V vs: SCE:
`
`ð5Þ
`
`The protons so generated will be driven to the catholyte
`electro-statically and partially supplement the protons
`consumption during the synthesis of H2O2.
`Polarization curves reflect transient current response
`with respect to cathodic potential ðEcÞ applied. Results
`in Fig. 3 indicate that at Eco0:15 V,
`the current
`rapidly with increasing Ec:
`density (i)
`increases
`However, a ‘‘plateau’’ appears in the range of 0.15
`to 0.5 V. This ‘‘plateau’’ represents the limiting current
`region for the electro-generation of H2O2 (Reaction 2).
`When the Ec continues to increase above 0.5 V, the i
`quickly rises again. It implies a significant reduction of
`H2O2 to H2O (Reaction 3) and an enhanced H2
`evolution (Reaction 4). Pure oxygen gas provides a
`higher DO concentration than air, thereby yielding a
`higher i: Fig. 3 also shows that the initial equilibrium
`potential, E0; is 0.075 V at i ¼ 0: If a constant potential
`of 0.5 V is applied,
`the initial overpotential
`is
`calculated as 0.425 V. As electrolysis proceeds, the
`overpotential becomes more negative since the E0
`
`Fig. 3. Polarization curves of pure oxygen and air sparging.
`Experimental
`conditions:
`completely mixing,
`sweeping
`rate=33.3 mV/s; pH=2; T ¼ 231C; QO2 ¼ 8:2  102 mol/min;
`Qair ¼ 8:2  102 mol/min;
`ionic strength=0.05 M NaClO4;
`long-finger plate cathode.
`
`where D represents the diffusion coefficient of oxygen
`(m2/s) and d is the thickness of diffusion layer ðmÞ: Fig. 3
`shows that the maximum limiting current is approxi-
`mately located in the range of 0.4–0.5 V ðEcÞ:
`Current efficiency ðZÞ; defined as the ratio of the
`electricity consumed by the electrode reaction of interest
`over the total electricity passed through the circuit, can
`be calculated by Eq. (8):
`R
`Z ¼ nFCH2O2 V
`I dt
`
`100%;
`
`t 0
`
`ð8Þ
`
`where CH2O2 represents H2O2 concentration in bulk
`solution ðMÞ and V is the catholyte volume ðLÞ: By
`definition, the Z actually represents an overall current
`efficiency over a certain period of electrolysis time.
`The effect of cathodic potential was investigated from
`0.2–0.9 V ðEcÞ: Figs. 4a–c shows the time-dependent
`changes of H2O2 concentration, current density and
`current efficiency at various applied potentials, respec-
`tively. Results indicate that at 0:2p Ecp0:5 V, the
`H2O2 concentration increases linearly with reaction time
`(Fig. 4a). The slope represents a constant net generation
`rate of H2O2 ðgnÞ throughout the whole electrolysis
`course. Correspondingly, Fig. 4b shows that i stabilize
`quickly after the electrolysis is initiated. A steady-state
`condition is rapidly reached because a constant DO
`concentration is maintained in the solution. Fig. 4c
`indicates a slight decrease of Z during electrolysis
`because the reduction of H2O2 to H2O is gradually
`promoted by the accumulation of H2O2. Results further
`indicate that at Ec > 0:5 V, both gn and Z decrease
`notably with reaction time (Figs. 4a and c). It implies
`that a high Ec stresses the reduction of H2O2 as well as
`the evolution of H2. Fig. 4b shows that at 0.6 V, the i
`
`

`

`90
`
`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`Fig. 4. Generation of H2O2 at various applied cathodic potentials: (a) accumulated concentration; (b) current density; (c) current
`conditions: pH=2; T ¼ 231C;
`efficiency;
`(d)
`average
`current density
`and average
`current
`efficiency. Experimental
`QO2 ¼ 8:2  102 mol/min; ionic strength=0.05 M NaClO4; long-finger plate cathode.
`
`exhibits a slight increase with reaction time. Meanwhile,
`a remarkable increase of i is observed at 0.8 and 0.9 V.
`Though more electricity is consumed at high Ec; a
`higher fraction of which is wasted by side reactions. The
`maximum H2O2 concentration is obtained at 0.6 V, i.e.,
`79 mg/l after electrolysis for 2 h. However, based on Z; a
`constant potential of 0.5 V (vs. SCE) is the optimal
`i at all applied
`cathodic potential. The high initial
`potentials in Fig. 4b is caused by the pre-saturated DO
`concentration on the cathode surface. By plotting the
`average i and average Z vs. Ec (Fig. 4d), it is clear that
`the optimal cathodic potential is 0.5 V vs. SCE, and
`the corresponding i and Z are 6.4 A/m2 and 81%,
`respectively. The i obtained at the optimal potential, i.e.,
`6.4 A/m2, is called ‘‘the limiting current density’’.
`
`3.3. Effect of oxygen purity and mass flow rate
`
`Pure oxygen gas (99.6%) and air were used as the
`sources of DO. Both oxygen purity and mass flow rate
`
`affect the limiting current. The equilibrium DO con-
`centration is proportional to the oxygen partial pressure
`in the supply gas. For example, the equilibrium DO
`concentrations were measured as 8.3 and 39.3 mg/l when
`sparging air and pure oxygen at pH 2 and 0.05 M
`NaClO4, respectively. Fig. 5a shows the accumulation of
`H2O2 as a function of reaction time at various oxygen
`purities and mass flow rates. Results clearly indicate that
`gn increases with increasing oxygen flow rate until a rate
`of 8.2  102 mol/min is achieved. Further increase in
`the flow rate does not change the gn: Results also
`indicate that the gn is much smaller when air is sparged.
`Fig. 5b shows the effect of mass flow rate on the average
`i and average Z: In accordance with Fig. 5a, the average i
`reaches the maximum value of 6.4 A/m2 at the oxygen
`flow rate of 8.2  102 mol/min. Doubling the flow rate
`will not increase the i any more. It is seen that a rate of
`8.2  102 mol/min is adequate to maintain the highest
`steady-state DO concentration during electrolysis. Since
`the optimal potential was applied (Ec ¼ 0:5 V vs.
`
`

`

`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`91
`
`reached the steady-state condition (ca. 20 min). Fig. 6
`shows that a linear relationship is achieved between the
`limiting current and the effective surface area. The slope
`of 6.4 A/m2 is just the limiting current density. Results
`clearly indicate that
`the limiting current density is
`independent of the cathode geometry and surface area
`applied. It is obvious that increasing the surface area is a
`convenient way to raise the limiting current, and
`consequently, the net generation rate of H2O2. The
`km;
`is
`calculated
`as
`mass
`transfer
`coefficient,
`2.70  105 m/s using Eq. (6) and a C value of
`39.3 mg/l. Moreover, by assuming the oxygen diffusion
`coefficient ðDÞ as 2.0  109 m2/s [31], the thickness of
`the diffusion layer ðdÞ is calculated as about 74 mm using
`Eq. (7). Both km and d provide a detailed microscopic
`insight into the electrolytic synthesis of H2O2.
`
`3.5. Effect of pH
`
`From Eq. (2), it seems that a low pH is favorable for
`the electro-generation of H2O2
`since its
`synthesis
`consumes protons. However, a high proton concentra-
`tion may promote H2 evolution and reduce the current
`efficiency. Fig. 7 shows the effect of pH on gn and
`average Z in the constant current mode (i ¼ 6:4 A/m2).
`Results indicate that pH 2 is the optimal condition.
`Above pH 2, the gn decreases due to insufficient protons.
`Below pH 2, the gn decreases again due to enhanced H2
`evolution. Because pH 3 may stress the decomposition
`of H2O2 by trace metals as previously mentioned, the gn
`at pH 3 is a little smaller than at other pH values. Since
`the current was holding constant, the average Z curve
`exhibits a similar trend as the gn curve. The highest Z is
`84% at pH 2, and the lowest is 69% at pH 3.
`
`Fig. 6. Effect of cathode surface area on limiting current.
`conditions: Ec ¼ 0:5 V vs. SCE; pH=2;
`Experimental
`T ¼ 231C; QO2 ¼ 8:2  102 mol/min;
`ionic strength=0.05 M
`NaClO4.
`
`Fig. 5. Generation of H2O2 with pure oxygen or air at various
`mass flow rates: (a) accumulated concentration; (b) effect of
`mass flow rate on average current density and average current
`efficiency. Experimental conditions: Ec ¼ 0:5 V vs. SCE;
`pH=2; T ¼ 231C; ionic strength=0.05 M NaClO4; long-finger
`plate cathode.
`
`SCE), a high Z could be achieved for all flow rates, i.e.,
`80–90%. If air is sparged at a rate of 8.2  102 mol/min
`(or 1.7  102 mol O2/min), the average i decreases to
`2.1 A/m2, while the average Z slightly increases to 90%
`(Fig. 5b).
`
`3.4. Effect of cathode surface area
`
`i.e., plain
`Three graphite electrode configurations,
`plate,
`short- and long-finger plate were used to
`investigate the effect of cathode surface area. It should
`be pointed out that H2O2 was only generated at one side
`of the cathode that faces the applied electrical field. The
`effective surface areas were 271, 415 and 488 cm2 for the
`plain plate, short- and long-finger plate, respectively.
`The potential was controlled at 0.5 V vs. SCE, and the
`oxygen mass flow rate was 8.2  102 mol/min. The
`electrical current was measured after the electrolysis
`
`

`

`92
`
`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`Fig. 7. Effect of pH on net generation rate and average current
`i ¼ 6:4 A/m2; T ¼ 231C;
`efficiency. Experimental conditions:
`QO2 ¼ 8:2  102 mol/min;
`ionic strength=0.05 M NaClO4;
`long-finger plate cathode.
`
`temperature on net generation rate and
`Fig. 8. Effect of
`average current efficiency. Experimental conditions: i ¼ 6:4 A/
`m2; pH=2; QO2 ¼ 8:2  102 mol/min; ionic strength=0.05 M
`NaClO4; long-finger plate cathode.
`
`The effect of pH was also investigated using the
`constant potential mode at Ec ¼ 0:5 V vs. SCE (data
`not shown). Results indicate that both gn and i increases
`with decrease in pH, but the highest Z (81%) is still
`achieved at pH 2. Therefore, it is concluded that pH 2 is
`the optimal value for H2O2 generation in acidic
`solutions.
`
`3.6. Effects of temperature and supporting electrolyte
`
`Temperature exerts conflicting effects on H2O2 gen-
`eration. As temperature rises,
`the oxygen diffusion
`coefficient will increase, resulting in an increase of gn:
`However, increasing temperature will decrease the DO
`solubility and increase the H2O2 decomposition rate,
`thereby decreasing gn: If the current is held constant, the
`formation rate of H2O2 will not change appreciably, but
`the decomposition rate of H2O2 will
`increase as
`temperature rises. The effect of
`temperature was
`investigated from 131C to 331C using the constant
`current mode (i ¼ 6:4 A/m2). Fig. 8 demonstrates that gn
`slightly decreases with increasing temperature. The
`average Z shows a similar trend, decreasing from 92%
`at 131C to 81% at 331C. It is seen that the H2O2
`generation is favored at low temperatures.
`The effect of inert supporting electrolyte concentra-
`tion was also investigated using the constant current
`mode. Fig. 9 shows that
`in the region of NaClO4
`concentration less than 0.1 M, the total potential drop
`between cathode and anode ðDEÞ decreases notably with
`increasing NaClO4 concentration. In the >0.1 M region,
`the DE decreases slowly due to the decrease of ion
`activity coefficients. Since NaClO4 does not participate
`the gn remains almost
`in electrochemical reactions,
`constant in the concentration range studied. Do and
`Chen [24]
`investigated the oxidative degradation of
`
`Fig. 9. Effect of NaClO4 concentration on total potential drop
`and net generation rate. Experimental conditions: i ¼ 6:4 A/m2;
`pH=2; T ¼ 231C; QO2 ¼ 8:2  102 mol/min; long-finger plate
`cathode.
`
`formaldehyde with electro-generated H2O2 and reported
`that the formaldehyde degradation was not affected by
`the concentration of supporting electrolyte, Na2SO4, in
`the range of 0–0.5 M. The results indicate that the
`background ionic strength in wastewater may be directly
`used as supporting electrolyte for on-site H2O2 genera-
`tion.
`It has been reported that chloride or bromide can
`promote H2O2 generation [32,33]. Since the SCE
`reference was filled with saturated KCl solution, the
`diffusion of chloride from the SCE into catholyte might
`enhance H2O2 generation and skew the experimental
`data. The experimental
`results
`from the constant
`potential mode (Ec ¼ 0:5 V vs. SCE, with SCE inserted
`in the catholyte) and from the constant current mode
`(i ¼ 6:4 A/m2, without SCE) are compared. During 2 h
`
`

`

`Z. Qiang et al. / Water Research 36 (2002) 85–94
`
`93
`
`of electrolysis, the total electricity consumed is very
`close, i.e., 2,257 C for the constant potential mode versus
`2,232 C for the constant current mode. The gn and Z are
`0.64 mg/l min and 81% (Figs. 4a and d), and 0.68 mg/
`l min and 84% (Fig. 7), correspondingly. It is seen that
`the presence of the SCE in the catholyte does not
`promote H2O2 generation. The diffusion of chloride is
`negligible, mainly due to the large reactor volume with a
`relatively short electrolysis time.
`
`4. Conclusions
`
`The major objective of this study is to improve the
`Faradic current efficiency of H2O2 generation in acidic
`solutions containing dilute supporting electrolyte by
`optimizing the operational parameters. Based on the
`experimental results presented above,
`the following
`conclusions are drawn:
`
`3.7. Potential profile and energy consumption
`
`Fig. 10 shows the potential profile in the electrolyzer
`at Ec ¼ 0:5 V vs. SCE. Results indicate that the DE
`consists of three parts, i.e., the potential drops at the
`electrode–solution interfaces caused by activation polar-
`ization and concentration polarization of DO, and the
`IR drop in the electrolyte solution. In the solution,
`Ohm’s law is well obeyed since the electrolyte concen-
`tration remains almost constant. The electrical resis-
`tance of the cation exchange membrane is negligible,
`which means a prompt migration of protons from
`anolyte to catholyte. A constant anolyte pH was
`observed during the course of electrolysis although
`protons were being continuously generated at the anode.
`Under the experimental conditions stated in Fig. 10,
`the unit energy consumption for H2O2 generation is
`calculated as 7.8 kW h/kg H2O2.
`
`* Significant self-decomposition of H2O2 is observed
`only at high pH (>9) and elevated temperatures
`(>231C).
`* The optimal conditions for H2O2 generation are
`cathodic potential of 0.5 V (vs. SCE), oxygen mass
`flow rate of 8.2  102 mol/min, and pH 2. Under the
`optimal conditions, the average current density and
`average current efficiency are 6.4 A/m2 and 81%,
`respectively. When air is used, the average current
`density decreases to 2.1 A/m2, while the average
`current efficiency slightly increases to 90%.
`i.e., 6.4 A/m2,
`* The limiting current density,
`is
`independent of the cathode geometry and surface
`area applied.
`* H2O2 generation is
`tures.
`* In the concentratio

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket