Journal of Harbin Institute of Technology  2017, Vol. 24 Issue (4): 9-17  DOI: 10.11916/j.issn.1005-9113.16116
0

Citation 

Shaoying Qi, Jinfeng Lu. Mathematical Advancement of General Chlorine Bulk Decay Model[J]. Journal of Harbin Institute of Technology, 2017, 24(4): 9-17. DOI: 10.11916/j.issn.1005-9113.16116.

Fund

Sponsored by the National Key Research and Development Program of China (Grant No.2016YFC0400707), National Science and Technology Major Project of the Ministry of Science and Technology of China (Grant No.2015ZX07203-007), National Natural Science Foundation of China (Grant No.31601834) and Natural Science Foundation of Tianjin (Grant No.14JCYBJC22900)

Corresponding author

Shaoying Qi, E-mail:sqi@illinois.edu

Article history

Received: 2016-05-18
Mathematical Advancement of General Chlorine Bulk Decay Model
Shaoying Qi1, Jinfeng Lu2,3     
1. Department of Civil & Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, USA;
2. College of Environmental Science and Engineering, Nankai University, Tianjin 300071, China;
3. Key Laboratory of Pollution Processes and Environmental Criteria(Nankai University), Ministry of Education, Tianjin 300071, China
Abstract: An accurate chlorine bulk decay model is needed to ensure that potable water meets the microbial and the chemical safeties at the treatment plant and throughout the distribution. Among the mathematical models available, the general second-order chlorine bulk decay model (GBDM) is the most fundamentally sound. Application of the GBDM, however, has been hindered by its numerous fictive parameters and lack of an analytical solution. This theoretical work removes the two obstacles. The GBDM is solved through transformation and integration. The analytical solution provides deep insights into the GBDM and facilitates the parameterization and sensitivity analysis. The background natural organic matter (NOM) is characterized with the probabilistic distribution of functional groups. It reveals that the mean of the function group distribution is correlated with the initial chlorine decay rate coefficient (κ0). A simple formula is developed to determine κ0 directly from the initial chlorine decay. The theoretical treatment reduces the fictive parameters to a minimum. For the common lognormal distribution, the GBDM needs only three parameters, well defined as initial chlorine demand X0, median rate coefficient km, and heterogeneity index σ. For more complicated scenarios, composite distributions are constructed through superposition of individual distributions. A highlighted example is to predict chlorine decay in blends of different waters. With the theoretical and mathematical advancement here, the GBDM can be applied effectively to any reactive background matter in any reaction systems.
Key words: chlorine residual     potable water     natural organic matter     decay model    
1 Introduction

Chlorine disinfection is routinely used by water utilities in P. R. China, the United States, and many other countries[1-3]. The amount of chlorine applied to a disinfection process must be sufficient to compensate the reaction loss and maintain the required residual over the entire disinfection time. The chlorine residual, on the other hand, cannot be too high to produce excess disinfection byproducts (DBPs) and/or make the water taste and smell like chlorine. An upper limit is typically set at below the maximum residual disinfectant level (MRDL) of 4 mg/L Cl2[4]. To prevent the microbial regrowth, potable water is also required to maintain the chlorine residual throughout distribution. The RMDL (required minimum disinfectant level), for example, is 0.2 mg/L Cl2[4]. At the treatment plant, the bulk decay or reaction with the substances in bulk water dominates the chlorine consumption. Once in the distribution system, the bulk decay continues while the wall decay also occurs (Note: the wall decay, which is caused by "the pipe walls and the biofilm and corrosion products that adhere to them"[5], is beyond the scope of this study). The bulk decay can be measured independently while the wall decay cannot or has to be determined through subtracting the bulk decay from the total decay. An accurate chlorine bulk decay model will serve to ensure that the potable water meets the microbial and the chemical safeties at the treatment plant and throughout the distribution.

Among the mathematical models available, the general second-order chlorine bulk decay model (GBDM) is most fundamentally sound[3, 6-10]. The GBDM is a multiple channel reaction model that can handle all the reactive substances in water. From the GBDM, various usable models are developed either as direct simplifications (by retaining just one, two or three reactants) or as semi-theoretical variable rate coefficient (VRC) models. Those usable models are accompanied by strong experimental evidences[3, 6-11]. The dispute over which one is better, however, has not been settled. Most noticeably, some researchers[6-8, 10] recommend the two-reactant (2R) model while others[3, 9] question the adequacy of two reactants and favor the semi-theoretical VRC models. To resolve the difference over the possible scenarios, one will need the GBDM and Monte Carlo simulations.

Blends of different waters are increasingly used in area/countries where water sources are scarce. Fisher et al.[12] recently formulated "a dynamic model of chlorine decay in a blend of two sources" by "linking the two-reactant (2R) models of individual sources". The resulting blend decay model (BDM) has four (4) notional (fictive) reactants. Fisher et al.[12] demonstrated that their BDM with input parameters determined for individual sources (before blending) can predict the chlorine decay in blends of surface waters, desalinated water, and ground waters. This newly developed BDM is a special case of the GBDM and can be solved analytically as well.

The general second-order chlorine bulk decay model (GBDM) provides a theoretical basis for various usable models[3, 6-10]. The real application of the GBDM, i.e., treating the complex background matter truly as a mixture of many reactive compounds, however, has been hindered by its numerous fictive parameters and lack of an analytical solution. This theoretical work intends to remove the two obstacles and advance the understanding and application of the GBDM. The comprehensive study is organized into three sections. In section 2, we build the GBDM rationally from the individual reactions, solve the GBDM through transformation and integration, and treat the leading usable models as special cases. In section 3, we characterize the reaction heterogeneity with the probabilistic distribution, apply the theory to various scenarios, and revise the analytical solution accordingly. In section 4, we explore the GBDM in various equivalent forms, unify the interpretation of usable models, and establish the basic parameters for assessing/ranking the reactive background matter.

2 GBDM Development

Chlorine decay in potable water is primarily caused by the background natural organic matter (NOM). With the current state of knowledge, it is not infeasible to determine and quantify the NOM compounds individually. The complex NOM is thus treated as a mixture of N number fictive compounds. Each fictive compound is identified with elementary rate coefficient ki and initial concentration Ci, 0. The reaction with chlorine, which is redox in nature, is assumed to be bimolecular and irreversible. By the law of mass action, the rate of reaction is written:

$ {{r}_{i}}=-{{k}_{i}}{{C}_{\text{Cl}}}{{C}_{i}}\left( i=1\ \text{to}\ N \right) $ (1)

where Ci and CCl are the concentrations of compound i and chlorine (Cl2), respectively. The chlorine equivalency, which is a known 1:1 stoichiometric ratio (just like the electron equivalency), is selected and recommended as the measure of all the reactants. The overall reaction rate is obtained as:

$ {{r}_{\text{b}}}=\sum\limits_{i=1}^{N}{{{r}_{i}}}=-{{C}_{\text{Cl}}}\sum\limits_{i=1}^{N}{{{k}_{i}}{{C}_{i}}} $ (2)

The rate of chlorine decay is equal to the overall reaction rate given in Eq.(2). Eqs.(1) and (2) complete the basic rate equations for the general second-order chlorine bulk decay model (GBDM). Evidently, the chlorine decay, which occurs through multiple reaction channels, is modeled with all parallel bimolecular reactions. Any sequential chlorine reaction is assumed to be insignificant or approximated by the parallel reactions. Mathematical justification for representing serial processes with parallel reactions may be found in Forney and Rothman[13] and Bolker et al.[14]. Eqs.(1) and (2) can be extended directly to any mixture of compounds regardless of whether they are real, fictive, known or unknown.

Incorporating Eqs.(1) and (2) into the mass balance over a batch or plug-flow reactor results in a form of the GBDM often seen in the literature[3, 6-9]

$ \frac{\text{d}{{C}_{i}}}{\text{d}t}=-{{k}_{i}}{{C}_{\text{Cl}}}{{C}_{i}}\left( i=1\ \text{to}\ N \right) $ (3)
$ \frac{\text{d}{{C}_{\text{Cl}}}}{\text{d}t}=\sum\limits_{i=1}^{N}{\frac{\text{d}{{C}_{i}}}{\text{d}t}}=-{{C}_{\text{Cl}}}\sum\limits_{i=1}^{N}{{{k}_{i}}{{C}_{i}}} $ (4)
$ {{C}_{\text{Cl}}}\left( t=0 \right)={{C}_{\text{Cl,0}}}\ \text{and}\ {{C}_{i}}\left( t=0 \right)={{C}_{i\text{,0}}} $ (5)

where CCl, 0 and Ci, 0 are the initial concentrations of chlorine and compound i, respectively. Eq.(4) can be integrated and replaced by:

$ {{C}_{\text{Cl,0}}}-{{C}_{\text{Cl}}}=\sum\limits_{i=1}^{N}{\left( {{C}_{i\text{,0}}}-{{C}_{i}} \right)}={{X}_{0}}-X $ (6)
$ {{X}_{0}}=\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}} $ (7)
$ X=\sum\limits_{i=1}^{N}{{{C}_{i}}} $ (8)

where X0 is the initial chlorine demand and X is the (remaining) chlorine demand at reaction time t. With a sufficiently large dose of chlorine: CCl, 0> X0, X(t→∞)= 0 and Eq.(6) reduces to

$ {{X}_{0}}={{C}_{\text{Cl,0}}}-{{C}_{\text{Cl}}}\left( t\to \infty \right) $ (9)

where CCl(t→∞)> 0. The measurability of X0 and X, which are the exact amounts of reactive background matter present initially and at time t, further strengths the use of chlorine equivalency with the GBDM. It is also worthwhile to recognize that other measures/units being used in the literature may cause unnecessary complication and/or theoretical inconsistency. For example, the total organic carbon (TOC) is stoichiometrically irresponsible (and might even remain unchanged) during chlorination while the molar ratio is unknown and likely to vary among the NOM compounds.

Eq.(3) is an autonomous system of first order ordinary differential equations (ODEs). Dividing any two of the ODEs with each other eliminates both variables t and CCl(t)

$ \frac{\text{d}{{C}_{i}}}{\text{d}{{C}_{j}}}=\frac{{{k}_{i}}}{{{k}_{j}}}\frac{{{C}_{i}}}{{{C}_{j}}}\left( i=1\ \text{to}\ N\ \text{for}\ \text{any}\ j=1\ \text{to}\ N \right) $ (10)

Eq.(10) is integrated:

$ {{C}_{i}}={{C}_{i\text{,0}}}{{\left[ {{C}_{j}}/{{C}_{j\text{,0}}} \right]}^{\left( {{k}_{i}}/{{k}_{j}} \right)}} $ (11)

Subsequently, substituting Eq.(11) into Eqs.(8) and (6) gives:

$ X=\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}{{\left[ {{C}_{j}}/{{C}_{j\text{,0}}} \right]}^{\left( {{k}_{i}}/{{k}_{j}} \right)}}} $ (12)
$ {{C}_{\text{Cl}}}={{C}_{\text{Cl,0}}}-{{X}_{0}}+\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}{{\left[ {{C}_{j}}/{{C}_{j\text{,0}}} \right]}^{\left( {{k}_{i}}/{{k}_{j}} \right)}}} $ (13)

To relate Cj/Cj, 0 to reaction time t, Eq.(13) is substituted back into Eq.(3). The resulting ODE is integrated:

$ \int_{1}^{\left( {{C}_{j}}/{{C}_{j\text{,0}}} \right)}{\frac{1/x}{{{C}_{\text{Cl,0}}}-{{X}_{0}}+\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}{{x}^{\left( {{k}_{i}}/{{k}_{j}} \right)}}}}\text{d}x}=-{{k}_{j}}t $ (14)

Although it cannot be expressed in elementary functions, the integral can be evaluated to any desired level of accuracy using any of the quadrature formulas[15].

Either chlorine residual CCl(t) or reaction time t(CCl) can be predicted with Eq.(13) or Eq.(14). The two equations, which are linked through intermediate variable Cj/Cj, 0 only, can be solved in sequence. Let

$ y=\frac{{{C}_{j}}}{{{C}_{j\text{,0}}}} $ (15)

To determine time t(CCl) for CCl, 0 to reduce to CCl, one may first solve Eq.(13) using the Newton-Raphson method

$ {{y}_{+}}={{y}_{c}}-\frac{{{C}_{\text{Cl,0}}}-{{C}_{\text{Cl}}}-{{X}_{0}}+\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}y_{c}^{\left( {{k}_{i}}/{{k}_{j}} \right)}}}{\sum\limits_{i=1}^{N}{{{C}_{i\text{,0}}}\left( {{k}_{i}}/{{k}_{j}} \right)y_{c}^{\left( {{k}_{i}}/{{k}_{j}}-1 \right)}}} $ (16)

where yc is the current estimate and y+ is the new estimate for y (0≤y≤1). Once a desirable value for y is obtained, t(CCl) is computed directly using Eq.(14). Likewise, to predict residual CCl(t) after time t, one may first solve Eq.(14) using the Newton-Raphson method

$ \begin{array}{l} {y_ + } = {y_c} - {y_c}\left[ {{C_{{\rm{Cl,0}}}} - {X_0} + \sum\limits_{i = 1}^N {{C_{i{\rm{,0}}}}y_c^{\left( {{k_i}/{k_j}} \right)}} } \right]\left[ {{k_j}t + } \right.\\ \;\;\;\;\;\;\;\left. {\int_1^{{y_c}} {\frac{{1/x}}{{{C_{{\rm{Cl,0}}}} - {X_0} + \sum\limits_{i = 1}^N {{C_{i{\rm{,0}}}}{x^{\left( {{k_i}/{k_j}} \right)}}} }}{\rm{d}}x} } \right] \end{array} $ (17)

and then compute CCl(t) using Eq.(13).

The well-known two-reactant (2R) model is developed by retaining just two terms of Eqs.(3) and (4)[6-8, 16]. The expectation is that one reactant would control the initial fast decay while the other reactant dominates the slow decay that follows. The two reactants are named as the fast (F) and the slow (S) reducing agents. From Eqs.(13) and (14) with N= 2, we obtain an analytical solution for the 2R model:

$ {C_{{\rm{Cl}}}} = {C_{{\rm{Cl,0}}}} - \left( {{F_0} + {S_0}} \right) + F + {S_0}{\left( {F/{F_0}} \right)^{\left( {{k_{\rm{S}}}/{k_{\rm{F}}}} \right)}} $ (18)
$ \int_1^{\left( {F/{F_0}} \right)} {\frac{{1/x}}{{{C_{{\rm{Cl,0}}}} - \left( {{F_0} + {S_0}} \right) + {F_0}x + {S_0}{x^{\left( {{k_{\rm{S}}}/{k_{\rm{F}}}} \right)}}}}{\rm{d}}x} = - {k_{\rm{F}}}t $ (19)

where F0 and S0 are the initial concentrations, F and S are the concentrations at time t, and kF and kS are the rate coefficients. Through an algebraic manipulation, Eqs.(18) and (19) can be shown to be the same as the solution given by Kohpaei and Sathasivan[7]. The 2R model[6-8, 16], however, is currently formulated with two deficiencies: 1) retaining just two terms of Eqs.(3) and (4)[6-8, 16] detaches the model from the GBDM; and 2) the F and the S reducing agents are ambiguously defined. In section 4, we will make the necessary clarification and develop/interpret the 2R model more coherently.

Fisher et al.[12] recently formulated "a dynamic model of chlorine decay in a blend of two sources" "by linking the two-reactant (2R) models of individual sources through weighting the initial concentrations" by dilution factors. This blend decay model (BDM) is also a special case of the general second-order chlorine bulk decay model (GBDM). Let's consider water Sources 1 and 2 with parameters listed in Table 1. The two waters are blended at volume ratio a/(1-a), where a and (1-a) are the volume fractions of Sources 1 and 2. The blend or composite water has four (4) fictive compounds at initial concentrations aF1, 0, aS1, 0, (1-a)F2, 0 and (1-a)S2, 0, respectively. From Eqs.(13) and (14) with N= 4, we obtain an analytical solution for the BDM:

Table 1 BDM parameters from individual water sources[12]

$ \begin{array}{l} {C_{{\rm{Cl}}}} = {C_{{\rm{Cl,0}}}} - {X_0} + a\left[ {{F_1} + {S_{1,0}}{{\left( {{F_1}/{F_{1,0}}} \right)}^{\left( {{k_{{\rm{S,1}}}}/{k_{{\rm{F,1}}}}} \right)}}} \right] + \\ \;\;\;\;\;\;\;\;\left( {1 - a} \right)\left[ {{F_{2,0}}{{\left( {{F_1}/{F_{1,0}}} \right)}^{\left( {{k_{{\rm{F,2}}}}/{k_{{\rm{F,1}}}}} \right)}} + } \right.\\ \;\;\;\;\;\;\;\;\left. {{S_{2,0}}{{\left( {{F_1}/{F_{1,0}}} \right)}^{\left( {{k_{{\rm{S,2}}}}/{k_{{\rm{F,1}}}}} \right)}}} \right] \end{array} $ (20)
$ \int_1^{\left( {{F_1}/{F_{1,0}}} \right)} {\frac{{1/x}}{{{C_{{\rm{Cl,0}}}} - {X_0} + a\left[ {{F_{1,0}}x + {S_{1,0}}{x^{\left( {{k_{{\rm{S,1}}}}/{k_{{\rm{F,1}}}}} \right)}}} \right] + \left( {1 - a} \right)\left( {{F_{2,0}}{x^{\left( {{k_{{\rm{F,2}}}}/{k_{{\rm{F,1}}}}} \right)}}} \right) + {S_{2,0}}{x^{\left( {{k_{{\rm{S,2}}}}/{k_{{\rm{F,1}}}}} \right)}}}}{\rm{d}}x} = - {k_{{\rm{F,1}}}}t $ (21)

where X0 is the initial chlorine demand of the blend

$ {X_0} = a\left( {{F_{1,0}} + {S_{1,0}}} \right) + \left( {1 - a} \right)\left( {{F_{2,0}} + {S_{2,0}}} \right) $ (22)

The general second-order chlorine bulk decay model (GBDM) rationally treats the complex natural organic matter (NOM) with a large (N) number of fictive compounds. The number of fictive parameters grows rapidly at rate (2N+1) (Note: N is often prescribed but itself is a fictive parameter). For example, the simple 2R model of Eqs.(18) and (19) has four parameters while the BDM of Eqs.(20) and (21) has eight parameters. To effectively manage the fictive parameters, a theoretical treatment below is presented.

3 Distribution of Rate Coefficients

Natural organic matter (NOM), formed from biomass degradation through many random reactions and channels, is physically and chemically heterogeneous[17]. The chemical heterogeneity, although discrete in reality, may be systematically characterized with a continuous distribution. Noticeable examples are the NOM molecular size distribution[18-19]; the proton and the metal bindings by NOM[17, 20-21]; and the adsorption of NOM by activated carbon[22-23]. The lognormal distribution, either by itself or as a building block, describes the NOM heterogeneity well. This theoretical treatment is actually quite general and classical, owing to Pauling et al.[24], Karush and Sonenberg[25], and others[26]. Pauling et al.[24] first applied the lognormal distribution to the intramolecular heterogeneity (functional group variation within same macromolecules). Karush and Sonenberg[25] later on showed that the same approach was applicable to the intermolecular heterogeneity (functional group variation associated with different molecules).

Reaction heterogeneity of NOM can be both intramolecular and intermolecular. The general second-order chlorine bulk decay model (GBDM) makes no differentiation between the two while treating the NOM with N number fictive compounds. The fictive compounds are identified with elementary reaction rate coefficients [k0, k1, …, ki, …, kN] complemented by initial concentrations [C0, 0, C1, 0, …, Ci, 0, …, CN, 0], where Ci, 0 matches ki only and ki-1 < ki. An elementary rate coefficient is associated with the functional groups that have the same reaction activation energy (Note: the same functional groups may not necessarily belong to one particular compound). Semi-infinite range [0, ∞) is prescribed for ki (i=1 to N). This ensures that all the possible rate coefficients are included. A completely-inert compound has no chlorine demand: C0, 0= 0 for k0= 0; and an instantaneously-reactive compound is unlikely present: CN, 0= 0 as kN→∞. Let index i vary randomly between points 1 and N. Then, ki is realized as a discrete random variable with Ci, 0 as its occurring frequency. The probability for ki to assume a value from set [k0, k1, …, ki, …, kN] is:

$ P\left( {{k_i}} \right) = \frac{{{C_{i{\rm{,0}}}}}}{{\sum\limits_{j = 1}^N {{C_{j{\rm{,0}}}}} }} = \frac{{{C_{i{\rm{,0}}}}}}{{{X_0}}} $ (23)

Varying with discrete random ki (i=1 to N), P(ki) is a probability mass function

$ P\left( {{k_i}} \right) = \int_{\left( {{k_i} - \Delta k/2} \right)}^{\left( {{k_i} + \Delta k/2} \right)} {f\left( k \right){\rm{d}}k} $ (24)

where f(k) is the probability density function. Provided that difference Δk= (ki -ki-1) is small,

$ \int_{\left( {{k_i} - \Delta k/2} \right)}^{\left( {{k_i} + \Delta k/2} \right)} {f\left( k \right){\rm{d}}k} = \frac{{{\rm{d}}\mathit{\Gamma }}}{{{\rm{d}}k}}\left| {_{k = {k_i}}{\rm{d}}k} \right. = f\left( {k = {k_i}} \right){\rm{d}}k $ (25)

where F(k) is the cumulative distribution function. Finally, combining Eqs.(23), (24) and (25) results in

$ \begin{array}{l} \frac{{{C_{i{\rm{,0}}}}}}{{{X_0}}} = F\left( {k = {k_i} + \Delta k/2} \right) - F\left( {k = {k_i} - \Delta k/2} \right) = \\ \;\;\;\;\;\;\;\;f\left( {k = {k_i}} \right){\rm{d}}k \end{array} $ (26)

By matching the probabilities, the discrete distribution of ki (i= 1 to N) is coherently transformed into a continuous distribution of k and vice versa. The theoretical treatment opens the door for the GBDM to work with a family of probability distributions and reduces the fictive parameters to a minimum.

Chemical heterogeneity of NOM or a class of compounds may follow the law of lognormal distribution[13, 17, 19-20, 22-25, 27]. Eq.(26) then becomes:

$ {C_{i{\rm{,0}}}} = {X_0}\frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}\frac{1}{{{k_i}}}{{\rm{e}}^{ - {{\left[ {\left( {\ln {k_i} - \ln {k_{\rm{m}}}\sigma \sqrt 2 } \right)/} \right]}^2}}}{\rm{d}}k $ (27)

where km is the median rate coefficient, and σ is the variance of lnk distribution. Following Pauling et al.[24], σ is referred to as the heterogeneity index. A lognormal distribution of rate coefficient k is skewed with many small k values and fewer large ones. The skewness increases with σ,

$ skew = \left( {{{\rm{e}}^{{\sigma ^2}}} + 2} \right)\sqrt {{{\rm{e}}^{{\sigma ^2}}} - 1} $ (28)

Substituting Eq.(27) into Eq.(11) gives:

$ {C_i} = {X_0}{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)^{\left( {{k_i}/{k_{\rm{m}}}} \right)}}\frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}\frac{1}{{{k_i}}}{{\rm{e}}^{ - {{\left[ {\ln \left( {{k_i}/{k_{\rm{m}}}} \right)/\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}k $ (29)

Subsequently, substituting Eq.(29) into Eqs.(12), (13) and (14) yields:

$ {X = {X_0}\frac{1}{{\sigma \sqrt {2\pi } }}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m}},{\rm{0}}}}} \right)}^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } } $ (30)
$ \begin{array}{l} {C_{{\rm{Cl}}}} = {C_{{\rm{Cl}},{\rm{0}}}} - {X_0}\left[ {1 - \frac{1}{{\sigma \sqrt {2\pi } }}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m}},{\rm{0}}}}} \right)}^\gamma } \cdot } } \right.\\ \left. {\;\;\;\;\;\;\;\;\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } \right] \end{array} $ (31)
$ \begin{array}{l} \int_1^{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)} {\frac{{1/x}}{{{C_{{\rm{Cl,0}}}} - {X_0}\left( {1 - \frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}\int_0^\infty {{x^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } } \right)}}{\rm{d}}x} = \\ \;\;\;\;\;\;\;\; - {k_{\rm{m}}}t \end{array} $ (32)

Eqs.(31) and (32) are coupled through intermediate variable Cm/Cm, 0 only. The two equations can thus be solved sequentially, i.e., for Cm(t)/Cm, 0 and then for CCl(t) (or for Cm(CCl)/Cm, 0 and then for t(CCl)), very much in the same way shown earlier for solving Eqs.(13) and (14).

Heterogeneity index σ lies somewhere in the range of 0 ≤ σ < ∞. As σ→ 0, the Gaussian $ \delta \left( \eta \right) = \frac{1}{{\sigma \sqrt {2\pi } }}{{\rm{e}}^{- {{[\eta /(\sigma \sqrt 2 )]}^2}}} $ becomes a Dirac delta function as δ(η)=0 for η≠0 and $ \int_{-\infty }^\infty {\delta \left( \eta \right){\rm{d}}} \eta = 1$. The integral in Eq.(32) approaches to x

$ \begin{array}{l} \mathop {\lim }\limits_{\sigma \to 0} \frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}\int_0^\infty {{x^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } = \mathop {\lim }\limits_{\sigma \to 0} \int_{ - \infty }^\infty {{x^{{{\rm{e}}^\eta }}} \cdot } \\ \;\;\;\;\;\frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}{{\rm{e}}^{ - {{\left[ {\eta /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\eta = x \end{array} $ (33)

Likewise, the integral in Eq.(31) converges to (Cm/Cm, 0) as σ→ 0. Eqs.(32) and (31) reduce to:

$ \int_1^{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)} {\frac{1}{{x\left[ {{C_{{\rm{Cl,0}}}} - {X_0}\left( {1 - x} \right)} \right]}}{\rm{d}}x} = - {k_{\rm{m}}}t $ (34)
$ {C_{{\rm{Cl}}}} = {C_{{\rm{Cl,0}}}} - {X_0}\left[ {1 - \left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)} \right] $ (35)

where Cm, 0= X0 and Cm=X satisfy the conservation of mass. Evaluating the integral yields:

$ {C_{{\rm{Cl}}}} = \frac{{{C_{{\rm{Cl,0}}}} - {X_0}}}{{1 - \left( {{X_0}/{C_{{\rm{Cl,0}}}}} \right){{\rm{e}}^{{k_{\rm{m}}}\left( {{X_0} - {C_{{\rm{Cl,0}}}}} \right)t}}}}\left( {{C_{{\rm{Cl,0}}}} \ne {X_0}} \right) $ (36)
$ {C_{\mathit{Cl}}} = \frac{{{C_{{\rm{Cl,0}}}}}}{{1 + {C_{{\rm{Cl,0}}}}{k_{\rm{m}}}t}}\left( {{C_{{\rm{Cl,0}}}} = {X_0}} \right) $ (37)

Eqs.(36) and (37) conform to the analytical solution of the second-order single reaction bulk decay model[1]. The use of lognormal distribution is therefore valid even when there is only one reactive compound. As σ→∞, $ \delta \left( k \right) = \frac{1}{{\sigma \sqrt {2\pi } }}\frac{1}{k}{{\rm{e}}^{{{[{\rm{ln}}\left( {k/{k_{\rm{m}}}} \right)/(\sigma \sqrt 2 )]}^2}}} $ is also a Dirac delta function since δ(k)= 0 for k≠ 0 and $ \int_0^\infty {\delta \left( k \right){\rm{d}}k = {\rm{ }}1} $. The corresponding integral approaches to 1

$ \mathop {\lim }\limits_{\sigma \to 0} \frac{1}{{\sigma \sqrt {2{\rm{\pi }}} }}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } = 1 $ (38)

Eqs.(30) and (31) then reduce to:

$ X\left( t \right) = {X_0}\;{\rm{and}}\;{C_{{\rm{Cl}}}}\left( t \right) = {C_{{\rm{Cl,0}}}} $ (39)

In summary, the water or background matter is homogeneously reactive if σ= 0, heterogeneously reactive when 0 < σ < ∞, and homogeneously inactive as σ→∞.

A more complicated or composite distribution of rate coefficients may occur when the water is a blend of several sources and/or has some known (organic and/orinorganic) reducing species. We may characterize each water source with rate coefficient distribution Eq.(27), treat the known reducing species as outliers, and combine the individual contributions through superposition. Let′s take a blend of two waters as an example. Source C has initial chlorine demand X0, 1 and rate coefficient distribution LN1(km, σ12) and Source B has initial chlorine demand X0, 2 and rate coefficient distribution LN2(βm, σ22). After blending Sources C and B at volume ratio a/(1-a), the blend or composite water has initial chlorine demand X0

$ {X_0} = a{X_{0,1}} + \left( {1 - a} \right){X_{0,2}} $ (40)

and a bimodal lognormal (BLN) distribution of rate coefficients (as a linear combination of LN1(km, σ12) and LN2(βm, σ22)). Applying Eqs.(11) and (27) yields:

$ {C_i} = {\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)^{\left( {{k_i}/{k_{\rm{m}}}} \right)}}\frac{{a{X_{0,1}}}}{{{\sigma _1}\sqrt {2{\rm{\pi }}} }}\frac{1}{{{k_i}}}{{\rm{e}}^{ - {{\left[ {\ln \left( {{k_i}/{k_{\rm{m}}}} \right)/\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}k $ (41)
$ {B_i} = {\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)^{\left( {\beta /{k_{\rm{m}}}} \right)}}\frac{{\left( {1 - a} \right){X_{0,2}}}}{{{\sigma _2}\sqrt {2{\rm{\pi }}} }}\frac{1}{{{\beta _i}}}{{\rm{e}}^{ - {{\left[ {\ln \left( {{\beta _i}/{\beta _{\rm{m}}}} \right)/\left( {{\sigma _2}\sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\beta $ (42)

where Cm/Cm, 0 is the relative concentration that matches rate coefficient km. After summation, the total remaining amounts are obtained:

$ {X_1} = \frac{{a{X_{0,1}}}}{{{\sigma _1}\sqrt {2{\rm{\pi }}} }}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _1}\sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } $ (43)
$ {X_2} = \frac{{\left( {1 - a} \right){X_{0,2}}}}{{{\sigma _2}\sqrt {2{\rm{\pi }}} }}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^{\left( {{\beta _{\rm{m}}}/{k_{\rm{m}}}} \right)\gamma }}\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _2}\sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } $ (44)

Finally, Eqs.(43) and (44) are substituted into Eqs.(13) and (14) to give:

$ {C_{{\rm{Cl}}}} = {C_{{\rm{Cl,0}}}} - {X_0} +\\ \int_0^\infty {\left[ {\frac{{a{X_{0,1}}}}{{{\sigma _1}\sqrt {2{\rm{\pi }}} }}{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _1}\sqrt 2 } \right)} \right]}^2}}} + \frac{{\left( {1 - a} \right){X_{0,2}}}}{{{\sigma _2}\sqrt {2{\rm{\pi }}} }}{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^{\left( {{\beta _{\rm{m}}}/{k_{\rm{m}}}} \right)\gamma }}{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _2}\sqrt 2 } \right)} \right]}^2}}}} \right]} \frac{1}{\gamma }{\rm{d}}\gamma $ (45)
$ - {k_{\rm{m}}}t = \int_1^{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)} {\frac{{1/x}}{{{C_{{\rm{Cl,0}}}} - {X_0} + \int_0^\infty {\left[ {\frac{{a{X_{0,1}}}}{{{\sigma _1}\sqrt {2{\rm{\pi }}} }}{x^\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _1}\sqrt 2 } \right)} \right]}^2}}} + \frac{{\left( {1 - a} \right){X_{0,2}}}}{{{\sigma _2}\sqrt {2{\rm{\pi }}} }}{x^{\left( {{\beta _{\rm{m}}}/{k_{\rm{m}}}} \right)\gamma }}{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {{\sigma _2}\sqrt 2 } \right)} \right]}^2}}}} \right]\frac{1}{\gamma }{\rm{d}}\gamma } }}{\rm{d}}x} $ (46)

The two equations, also linked through intermediate variable Cm/Cm, 0 only, can be solved sequentially for Cm(t)/Cm, 0 first and then for CCl(t) (or for Cm(CCl)/Cm, 0 and then for t(CCl)).

The general second-order chlorine bulk decay model (GBDM)originally has two basic equations: rate equations Eqs.(1) and (2). Here we recommend that rate coefficient distribution Eq.(26) or (27) be the third basic equation. Given that potable water is typically dominated by background natural organic matter (NOM), a rational approach to implement the GBDM is to adopt the common lognormal distribution Eq.(27), measure the chlorine decay in batch reactors, and calibrate the model using the analytical solution Eqs.(29) through (31). Once initial chlorine demand X0, median rate coefficient km, and heterogeneity index σ are known, the GBDM in Eqs.(1), (2), and (27) can be applied to any reaction systems, such as disinfection contactors, clear wells/storage tanks, and distribution networks.

4 Mean Reaction Rate Coefficient(s)

The general second-order chlorine bulk decay model (GBDM) given by Eqs.(3) and (4) can be rearranged into a form that explicitly matches the overall reaction [3, 9-10]

$ \frac{{{\rm{d}}{C_{{\rm{Cl}}}}}}{{{\rm{d}}t}} = \frac{{{\rm{d}}X}}{{{\rm{d}}t}} = - \kappa {C_{{\rm{Cl}}}}X $ (47)
$ \kappa = \frac{1}{X}\sum\limits_{i = 1}^N {{k_i}{C_i}} = \sum\limits_{i = 1}^N {\left( {{C_i}/\sum\limits_{j = 1}^N {{C_j}} } \right){k_i}} $ (48)

Rate coefficient κ maintains several physical significances while varying with reaction time t∈[0, ∞). Firstly, κ is the rate coefficient of the overall reaction and is equal to the rate coefficient of chlorine decay. Secondly or furthermore, κ is the mean reaction rate coefficient associated with the functional group distribution of background matter. By differentiating Eq.(48), Jonkergouw et al.[3]and Hua et al.[9] showed:

$ \frac{{{\rm{d}}\kappa }}{{{\rm{d}}t}} < 0\left( {{\rm{if}}\;N \ge 2} \right)\;{\rm{and}}\;\frac{{{\rm{d}}\kappa }}{{{\rm{d}}t}} = 0\left( {{\rm{if}}\;N = 1} \right) $ (49)

i.e., the multiple channel reactions (N≥2) cause the mean reaction rate coefficient to decrease monotonically. The conceptual and mathematical deduction coincides with a long standing observation that the chlorine decay in potable water has a variable or continuously-declining rate coefficient.

After substituting Eqs.(11) and (12) into Eq.(48), an analytical solution for the mean reaction rate coefficient is obtained:

$ \kappa = \frac{{\sum\limits_{i = 1}^N {{k_i}{C_{i,0}}{{\left( {{C_j}/{C_{j,0}}} \right)}^{\left( {{k_i}/{k_j}} \right)}}} }}{{\sum\limits_{i = 1}^N {{C_{i,0}}{{\left( {{C_j}/{C_{j,0}}} \right)}^{\left( {{k_i}/{k_j}} \right)}}} }} $ (50)

where Cj(t)/Cj, 0 is given by Eq.(14). The initial chlorine decay rate coefficient is equal to:

$ {\kappa _0} = \kappa \left( {t = 0} \right) = \frac{1}{{{X_0}}}\sum\limits_{i = 1}^N {{k_i}{C_{i,0}}} = \sum\limits_{i = 1}^N {\left( {{C_{i,0}}/\sum\limits_{j = 1}^N {{C_{j,0}}} } \right){k_i}} $ (51)

For the lognormal distribution of functional groups, Eqs.(29) and (30) are substituted into Eq.(48) directly:

$ \kappa = \frac{{{k_m}\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } }}{{\int_0^\infty {{{\left( {{C_{\rm{m}}}/{C_{{\rm{m,0}}}}} \right)}^\gamma }\frac{1}{\gamma }{{\rm{e}}^{ - {{\left[ {\ln \gamma /\left( {\sigma \sqrt 2 } \right)} \right]}^2}}}{\rm{d}}\gamma } }} $ (52)

where Cj(t)/Cj, 0 is given by Eq.(32). Initially (t=0),

$ {\kappa _0} = \kappa \left( {t = 0} \right) = {k_m}{{\rm{e}}^{\frac{{{\sigma ^2}}}{2}}} $ (53)

which shows that initial chlorine decay rate coefficient κ0 is directly correlated with median rate coefficient km and heterogeneous index σ. When σ =0, κ0= km, when 0 < σ < ∞, κ0> km, and as σ→∞, $ \mathop {\lim }\limits_{\sigma \to \infty } \left( {\frac{{{\kappa _0}}}{{{k_{\rm{m}}}}}} \right) \to \infty $ where both κ0 → 0 and km →0.

Significantly, initial chlorine decay rate coefficient κ0 is equal to the mean reaction rate coefficient associated with the functional group distribution of background water in original water. For any water or functional group distribution, κ0 can be determined directly from the initial chlorine decay

$ {\kappa _0} = \frac{1}{{{C_{{\rm{Cl,0}}}}{X_0}}}\left( { - \frac{{{\rm{d}}{C_{{\rm{Cl}}}}}}{{{\rm{d}}t}}\left| {_{t \to 0}} \right.} \right) = \frac{1}{{{C_{{\rm{Cl,0}}}}{X_0}}}\left( {\frac{{{C_{{\rm{Cl,0}}}} - {C_{{\rm{Cl}}}}\left( {\Delta t} \right)}}{{\Delta t}}} \right)\left| {_{t \to 0}} \right. $ (54)

where X0 is measurable according to Eq.(9). The conceptual connection and measurability makes κ0 a unique kinetic parameter for accessing/ranking the background NOM (which varies from water to water).

The reactive background matter collectively controls the chlorine decay through its mean reaction rate coefficient. This mechanism naturally extends to any fraction of the background matter. If the background matter is split into slow (S) and fast (F) reactive fractions, for example,

$ {k_{{\rm{ave,S}}}} = \frac{1}{m}\sum\limits_{i = 1}^m {{k_i}} \;{\rm{and}}\;{k_{{\rm{ave,F}}}} = \frac{1}{{\left( {N - m} \right)}}\sum\limits_{i = m + 1}^N {{k_i}} $ (55)
$ {S_0} = \sum\limits_{i = 1}^m {{C_{i,0}}} \;{\rm{and}}\;S = \sum\limits_{i = 1}^m {{C_i}} $ (56)
$ {F_0} = \sum\limits_{i = m + 1}^N {{C_{i,0}}} \;{\rm{and}}\;F = \sum\limits_{i = m + 1}^N {{C_i}} $ (57)

the general second-order chlorine bulk decay model (GBDM) takes on the form

$ \frac{{{\rm{d}}C}}{{{\rm{d}}t}} = \frac{{{\rm{d}}X}}{{{\rm{d}}t}} = \frac{{{\rm{d}}F}}{{{\rm{d}}t}} + \frac{{{\rm{d}}S}}{{{\rm{d}}t}} $ (58)
$ \frac{{{\rm{d}}F}}{{{\rm{d}}t}} = - {k_F}{C_{{\rm{Cl}}}}F $ (59)
$ \frac{{{\rm{d}}S}}{{{\rm{d}}t}} = - {k_S}{C_{{\rm{Cl}}}}S $ (60)
$ {k_S} = \frac{1}{S}\sum\limits_{i = 1}^m {{k_i}{C_i}} = \sum\limits_{i = 1}^m {\left( {{C_i}/\sum\limits_{j = 1}^m {{C_j}} } \right){k_i}} $ (61)
$ {k_F} = \frac{1}{F}\sum\limits_{i = m + 1}^N {{k_i}{C_i}} = \sum\limits_{i = m + 1}^N {\left( {{C_i}/\sum\limits_{j = m + 1}^N {{C_j}} } \right){k_i}} $ (62)

where elementary reaction rate coefficients ki-1 < ki, kave, F and kave, S are the arithmetic averages, kF(t) and kS(t) are the concentration (population) weighted means, and integer m may assume any value between 1 and (N-1). Given that ki, 1 < ki, kave, F > kave, S is absolutely true. However, kave, F > kave, S guarantees neither kF(t=0)> kS(t=0) nor that the F fraction controls the initial chlorine decay. At any reaction time t∈[0, ∞), the relative contribution of the two fractions is governed by:

$ \frac{{{\rm{d}}F}}{{{\rm{d}}S}} = \frac{{\left( {\frac{{{\rm{d}}F}}{{{\rm{d}}t}}} \right)}}{{\left( {\frac{{{\rm{d}}S}}{{{\rm{d}}t}}} \right)}} = \frac{{{k_F}F}}{{{k_S}S}} $ (63)

For the F fraction to dominate the initial chlorine decay, condition F0kF(t=0)>>S0kS(t=0) must be met. If F(t)kF(t) and S(t)kS(t) are comparable, both F and S fractions contribute significantly to the chlorine decay. When F(t) kF(t) < < S(t) kS(t), on the other hand, the S fraction takes over the control of chlorine decay. In all the circumstances, mean rate coefficients, kF and kS continuously decline with reaction time t. Only when kS and kF are invariable with t, Eqs.(55) through (62) reduce to the existing two-reactant (2R) model. Conceptually, though, kS and kF defined by Eqs.(61) and (62) are the mean rate coefficients of S and F fractions while kS and kF in the existing 2R model [6-8, 16] are the rate coefficients of single reactants. There are (N-1) number ways to split the background matter into two fractions. Also the background mater can be divided into a maximum of N number fractions. Therefore, the GBDM may take on numerous equivalent forms. Those equivalent forms are the most logical basis for simplifying or approximating the GBDM. This significant aspect has been unfortunately overlooked by the usable models[1, 6-8, 16] that retain just one, two or a few terms of Eqs.(1) and (2).

Jonkergouw et al.[3, 10] and Hua et al.[9] recently proposed the semi-theoretical variable rate coefficient (VRC) models. Their idea is to approximate the overall reaction rate coefficient (κ) with a function of chlorine demand (X/X0). Jonkergouw et al.[3] started with:

$ \frac{{{\rm{d}}\kappa }}{{{\rm{d}}t}} = {\alpha _J}{C_{{\rm{Cl}}}}\left( {{k_{\min }}\kappa - {\kappa ^2}} \right) $ (64)

where αJ>0 is an empirical parameter and kmin is the smallest rate coefficient. After dividing Eq.(64) with Eq.(47), the differential equation is solved with initial condition κ(t=0) = κ0

$ \begin{array}{l} \kappa = \left( {{\kappa _0} - {k_{\min }}} \right){\left( {X/{X_0}} \right)^{{\alpha _J}}} + {k_{\min }} = \left( {{\kappa _0} - {k_{\min }}} \right)X_0^{ - {\alpha _J}} \cdot \\ \;\;\;\;\;\;{\left( {{X_0} - {C_{{\rm{Cl,0}}}} + {C_{{\rm{Cl}}}}} \right)^{{\alpha _J}}} + {k_{\min }} \end{array} $ (65)

Now it is clear that the Jonkergouw et al.'s VRC model[3, 10] approximates κ with a power function. Substituting Eq.(65) into Eq.(47) gives:

$ \frac{{{\rm{d}}{C_{{\rm{Cl}}}}}}{{{\rm{d}}t}} = \frac{{{\rm{d}}X}}{{{\rm{d}}t}} = - \left[ {\left( {{\kappa _0} - {k_{\min }}} \right){{\left( {X/{X_0}} \right)}^{{\alpha _J}}} + {k_{\min }}} \right]{C_{{\rm{Cl}}}}X $ (66)

Eq.(66) is solved analytically

$ \int_{{C_{{\rm{Cl,0}}}}}^{{C_{{\rm{Cl}}}}} {\frac{{1/x}}{{\left\{ {\left( {{X_0} - {C_{{\rm{Cl,0}}}} + x} \right)\left[ {\left( {{\kappa _0} - {k_{\min }}} \right)X_0^{ - {\alpha _J}}{{\left( {{X_0} - {C_{{\rm{Cl,0}}}} + x} \right)}^{{\alpha _J}}} + {k_{\min }}} \right]} \right\}}}{\rm{d}}x} = - t $ (67)

In contrast, the Hua et al.'s VRCmodel[9] approximates κ directly with an exponential function:

$ \kappa = {\kappa _0}{{\rm{e}}^{ - \left( {{\beta _H}/{X_0}} \right)\left( {{C_{{\rm{Cl,0}}}} - {C_{{\rm{Cl}}}}} \right)}} = {\kappa _0}{{\rm{e}}^{ - {\beta _H}\left( {1 - X/{X_0}} \right)}} $ (68)

where βH>0 is an empirical parameter. Substituting Eq.(68) into Eq.(47) gives:

$ \frac{{{\rm{d}}{C_{{\rm{Cl}}}}}}{{{\rm{d}}t}} = \frac{{{\rm{d}}X}}{{{\rm{d}}t}} = - {\kappa _0}{{\rm{e}}^{ - {\beta _H}\left( {1 - X/{X_0}} \right)}}{C_{{\rm{Cl}}}}X $ (69)

Eq.(69) is solved analytically as well

$ \int_{{C_{{\rm{Cl,0}}}}}^{{C_{{\rm{Cl}}}}} {\frac{{1/x}}{{\left( {{X_0} - {C_{{\rm{Cl,0}}}} + x} \right){{\rm{e}}^{ - \left( {{\beta _H}/{X_0}} \right)\left( {{C_{{\rm{Cl,0}}}} - x} \right)}}}}{\rm{d}}x} = - {\kappa _0}t $ (70)
5 Conclusions

The general second-order chlorine bulk decay model (GBDM) is clarified based on the concept of fictive compounds, the chlorine equivalency, and the law of mass action. It turns out that the GBDM may take on numerous equivalent forms. The original or basic formulation accentuates the elementary reactions. All the other forms can be derived or interpreted through the mean reaction rate coefficient(s). A suite of analytical solutions are developed for batch and plug flow reactors. Leading usable models in the literature are treated as special cases and solved analytically as well. The analytical solutions provide deep insights into the GBDM and can also facilitate the parameterization and sensitivity analysis.

The background natural organic matter (NOM) is systematically characterized with the probabilistic distribution of functional groups. The theoretical treatment reduces the fictive parameters to a minimum. For the common lognormal distribution, the GBDM needs only three parameters, well defined as initial chlorine demand X0, median rate coefficient km, and heterogeneity index σ. The GBDM is valid over the entire range of σ= 0 to σ→∞ even though the log normal distribution is limited to 0 < σ < ∞. The NOM or background matter is homogeneously reactive when σ= 0; heterogeneously reactive when 0 < σ < ∞; and homogeneously inactive as σ→∞. For more complicated scenarios, composite distributions can be constructed through superposition of individual distributions. A highlighted example is to predict chlorine decay in blends of different waters.

The initial chlorine decay rate coefficient (κ0) is equal to the mean reaction rate coefficient associated with the functional group distribution of original background mater in water. A simple formula is developed to determine κ0 directly from the initial chlorine decay. The conceptual connection and measurability makes κ0 a unique parameter for accessing/ranking the background NOM. With the initial rate coefficient (κ0) and the initial chlorine demand (X0) measured, the number of fictive parameters are reduced by two, regardless of the functional group distribution. For the lognormal distribution, κ0 is correlated directly with median rate coefficient km and heterogeneous index σ. The distribution can be described by pairing either km and σ or κ0 and σ. It is natural for the GBDM to use km and σ as inputs but it is recommended that the κ0 be reported explicitly next to the km.

Equipped with the concept and formula of functional group distribution, the general second-order chlorine bulk decay model (GBDM) can be applied effectively to any reactive background matter. Future work will focus on characterizing various waters with the GBDM, quantifying the pH and the temperature effects, and integrating the GBDM into the disinfection process and the distribution network models.

References
[1] Liu W, Qi S. Modeling and verifying chlorine decay and chloroacetic acid formation in drinking water chlorination. Frontiers of Environmental Science & Engineering in China, 2010, 4(1): 65-72. (0)
[2] Kawamura S. Integrated Design and Operation of Water Treatment Facilities. NY: John Wiley & Sons, Inc, 2000. (0)
[3] Jonkergouw P M R, Khu S T, Savic D A, et al. A variable rate coefficient chlorine decay model. Environmental Science & Technology, 2009, 43(2): 408-414. (0)
[4] CFR, 40 CFR 141 National Primary Drinking Water Regulations. US Government Printing Office, 2012. (0)
[5] Fisher I, Kastl G, Sathasivan A. A suitable model of combined effects of temperature and initial condition on chlorine bulk decay in water distribution systems. Water Research, 2012, 46(10): 3293-3303. DOI:10.1016/j.watres.2012.03.017 (0)
[6] Fisher I, Kastl G, Sathasivan A. Evaluation of suitable chlorine bulk-decay models for water distribution systems. Water Research, 2011, 45(16): 4896-4908. DOI:10.1016/j.watres.2011.06.032 (0)
[7] Kohpaei A J, Sathasivan A. Chlorine decay prediction in bulk water using the parallel second order model: An analytical solution development. Chemical Engineering Journal, 2011, 171(1): 232-241. DOI:10.1016/j.cej.2011.03.034 (0)
[8] Liu B, Reckhow D A, Li Y. A two-site chlorine decay model for the combined effects of pH, water distribution temperature and in-home heating profiles using differential evolution. Water Research, 2014, 53: 47-57. DOI:10.1016/j.watres.2014.01.010 (0)
[9] Hua P, Vasyukova E, Uhl W. A variable reaction rate model for chlorine decay in drinking water due to the reaction with dissolved organic matter. Water Research, 2015, 75: 109-122. DOI:10.1016/j.watres.2015.01.037 (0)
[10] Yuan Y, Zhong D, Hou X, et al. Estabishment and performance of variable rate coefficient model. Journal of Harbin Institute of Technology, 2010, 42(10): 1559-1563. (0)
[11] Monteiro L S P P, Viegas R M C, Covas D I C, et al. Assessment of current models ability to describe chlorine decay and appraisal of water spectroscopic data as model inputs. Journal of Environmental Engineering, 2016. DOI:10.1061/(ASCE)EE.1943-7870.0001149 (0)
[12] Fisher I, Kastl G, Sathasivan A, et al. General model of chlorine decay in blends of surface waters, desalinated water, and groundwaters. Journal of Environmental Engineering, 2015, 141(12). DOI:10.1061/(ASCE)EE.1943-7870.0000980 (0)
[13] Forney D C, Rothman D H. Common structure in the heterogeneity of plant-matter decay. Journal of the Royal Society Interface, 2012, 9(74): 2255-2267. DOI:10.1098/rsif.2012.0122 (0)
[14] Bolker B M, Pacala S W, Parton W J. Linear analysis of soil decomposition: Insights from the century model. Ecological Applications, 1998, 8(2): 425-439. DOI:10.1890/1051-0761(1998)008[0425:LAOSDI]2.0.CO;2 (0)
[15] Tuna J J. Handbook of Numerical Calculations in Engineering. New York: McGraw-Hill Publishing Company, 1989. (0)
[16] Kastl G J, Fisher I H, Jegatheesan V. Evaluation of chlorine decay kinetics expressions for drinking water distribution systems modelling. Aqua (Oxford), 1999, 48(6): 219-226. (0)
[17] Tipping E. Cation Binding by Humic Substances. Cambridge, UK: Cambridge University Press, 2002. (0)
[18] Cabaniss S E, Zhou Q H, Maurice P A, et al. A log-normal distribution model for the molecular weight of aquatic fulvic acids. Environmental Science & Technology, 2000, 34(6): 1103-1109. (0)
[19] Dietrich L A S, McInnis D P, Bolster D, et al. Effect of polydispersity on natural organic matter transport. Water Research, 2013, 47(7): 2231-2240. DOI:10.1016/j.watres.2013.01.053 (0)
[20] Perdue E M, Lytle C R. Distribution model for binding of protons and metal-ions by humic substances. Environmental Science & Technology, 1983, 17(11): 654-660. (0)
[21] Orsetti S, Andrade E M, Molina F V. Modeling ion binding to humic substances: Elastic polyelectrolyte network model. Langmuir, 2010, 26(5): 3134-3144. DOI:10.1021/la903086s (0)
[22] Li F S, Yuasa A, Chiharada H, et al. Polydisperse adsorbability composition of several natural and synthetic organic matrices. Journal of Colloid and Interface Science, 2003, 265(2): 265-275. DOI:10.1016/S0021-9797(03)00526-5 (0)
[23] Matsui Y, Yuasa A, Li F S. Overall adsorption isotherm of natural organic matter. Journal of Environmental Engineering-Asce, 1998, 124(11): 1099-1107. DOI:10.1061/(ASCE)0733-9372(1998)124:11(1099) (0)
[24] Pauling L, Pressman D, Grossberg A L. The serological properties of simple substances. Ⅶ. A quantitative theory of the inhibition by haptens of the precipitation of heterogeneous antisera with antigens, and comparison with experimental results for polyhaptenic simple substances and for azopro-teins. Journal American Chemical Society, 1944, 66(5): 784-792. DOI:10.1021/ja01233a039 (0)
[25] Karush F, Sonenberg M. Interaction of homologous alkyl sulfates with bovine serum albumin. Journal American Chemical Society, 1949, 71(4): 1369-1376. DOI:10.1021/ja01172a065 (0)
[26] Sips R. On the structure of a catalyst surface. Journal of Chemical Physics, 1948, 16(5): 490-495. DOI:10.1063/1.1746922 (0)
[27] Limpert E, Stahel W A, Abbt M. Log-normal distributions across the sciences: Keys and clues. Bioscience, 2001, 51(5): 341-352. DOI:10.1641/0006-3568(2001)051[0341:LNDATS]2.0.CO;2 (0)