You are on page 1of 209

OXYGENTRANSFERSTUDIES IN THE

COMPLETELY
MIXEDACTIVATEDSLUDGEPROCESS

by

Richard Oliver Mines, Jr.

Dissertation submitted to the Faculty of the

Virginia Polytechnic Institute and State University

in partial fulfillment of the requirements for the degree of

DOCTOROF PHILOSOPHY

in

Civil Engineering

APPROVED:

J. H. Sherrard, Chairman

J.P. Wightman J. T. Novak

R. E. Benoit W.R. Knocke

July, 1983

Blacksburg, Virginia
DEDICATION

To my Parents

Mr. Richard O. Mines

and

Mrs. Dreama B. Mines

And to my wife Jane

For their love, support, and encouragement

Throughout my life.

ii
ACKNOWLEDGEMENTS

I would like to thank the following persons for their guidance and

support during my graduate study at Virginia Tech:

To my wife, Jane, for her love, encouragement and financial support

throughout my academic career at VPI&SU.

Dr. Joseph H. Sherrard for serving as my major thesis advisor but

more important his encouragement, concern and friendship which made it

possible to survive the rigors of the Ph.D. program.

Dr. John T. Novak for serving on my committee and for his

encouragement and suggestions that contributed to this research

project.

Dr. Robert E. Benoit, Dr. James P. Wightman and Dr. William R.

Knocke for serving on my committee and for their support in this

effort.

Dr. Clifford W. Randall and Dr. Richard D. Walker for providing

financial assistance for which this research project would have never

been made possible.

Andy Mitchell, Gary Hart, Jaidev Kunjur and Chris Brown for making

life in the laboratory bearable.

Dr. Calvert c. Churn for his friendship, encouragement and help

throughout this endeavor.

Mr. E.G. "Feets" Willard for his assistance in obtaining lab

equipment and his friendship.

iii
To my parents, Mr. and Mrs. Richard O. Mines and my brother and

sister, Barry and Angie, whose stedfast love and support helped me

through the perilous moments.

To my wife's parents, Mr. and Mrs. Galen D. Saul for their support,

emotionally and financially.

Dr. Donald K. Jamison, a once in a life time person who took a

personal interest in my personal and professional development.

Adil Godrej and Steve Fesko for teaching me how to use the CMS

computer terminal.

Dr. Cal Sawyer for inspiring me and introducing me to the field of

Environmental Engineering.

The many friends that I have made in my 3.5 years at Virginia Tech

which are too numerous to list.

Ms. Lois Rowsey for typing this manuscript and Ms. Pat Shorten for

inking the drawings.

And to the VMI which has helped me to fulfill one of the most

important goals in my life.


TABLE OF CONTENTS

DEDICATION• ••••••••••••••••••••••••••••••••••••••••••••••••••••••••• ii

ACKNOWLEDGEMENTS
•• ••••••••••••••••••••••••••••••••••••••••••••••••• iii

LIST OF TABLES
••••.•••.•...••••••••.•.••.•.••.••••.•.•••.•••••.•.•... x

LIST OF FIGURES
•••••••••..•••••••••••..••••••••••.••.••••••.•••.•••• xi

CHAPTER

I. INTRODUCTION
••••••.••••••••••••••••.••.••••••••••••••••••••••• !

NEED FOR STUDY• •••••••••••••••••••••••••••••••••••••••••••• 2

OBJECTIVES
••.•.•••••••••••.•••.•.••.•••••••••••.•••••.••••• 4

II. LITERATUREREVIEW• •••••••••••••••••••••••••••••••••••••••••••• 5

CONVENTIONAL
BIOKINETIC THEORY
••••••••••••••••••••••••••••• S

DUALSUBSTRATELIMITATIONS••••••••••••••••••••••••••••••••• 7

Models for Dual Substrate Limitations ••••••••••••••••••• 8

Dual Substrate Limitations in Biological Treatment ••••• 1O

Importance of Nitrification •••••••••••••••••••••••••••• 12

Environmental Conditions Needed ••••••••••••••••••••• 15

Oxygen requirements •••••••.••••••••••••••.••••••• 15

Temperature .••••••••••••••••••••••••••••••••••••• 17

pH and alkalinity •••••••••••••••••••••••••••••••• 17

Substrate and product inhibition ••••••••••••••••• 18

Kinetics and Biokinetic Constants ••••••••••••••••••• 2O

OXYGEN
TRANSFER
••••••••••••••••••••••••••••••••••••••••••• 21

Theory • ••••••••••••••••••••••••••.••••••..•..•••••••••• 23

V
TABLE OF CONTENTS(cont'd.)

Aerator Test Methods ••••••••••••••••••••••••••••••••••• 27

Alpha, Beta, and Theta ••••••••••••••••••••••••••••••••• 29

Factors Affecting Transfer ••••••••••••••••••••••••••••• 31

Testing Procedure Difficulties •••••••••••••••••••••• 31

Scaleup ••••••••••••••••••••••••••••••••••••••••••••• 33

Factors Affecting a ••·••••••••••··•••·•·•••·•••••••33

Factors Affecting B •••••••••••••••••••••••••••••••• 34

Manipulation of Data •••••••••••••••••••••••••••••••• 34

Oxygen Transfer in Biological Systems •••••••••••••••••• 36

SUMMARY ••••••••••••••••••••••••••••••••••••••••••••••••• 39

III. METHODS
ANDMATERIALS
•••••••••••••••••••••••••••••••••••••••• 4O

EXPERIMENTAL
APPROACH
••••••••••••••••••••••••••••••••••••• 4O

LABORATORY
STUDIES•••••••••••••••••••••••••••••••••••••••• 40

Laboratory Apparatus ••••••••••••••••••••••••••••••••••• 42

Synthetic Feed ••••••••••••••••••••••••••••••••••••••••• 45

Acclimation and Start-up ••••••••••••••••••••••••••••••• 5O

Daily Protoco~••••••••••·••••••••••••••••••••••••••••••Sl

Analytical Procedures •••••••••••••••••••••••••••••••••• 53

Chemical oxygen demand •••••••••••••••••••••••••••••• 53

Temperature ••••••••••••••••••••••••••••••••••••••••• 54

pH •••••••••• C, •••••••••••••••••••••••••••••••••••••• 54

Suspended solids •••••••••••••••••••••••••••••••••••• 54

vi
tABLE OF CONTENTS(cont'd.)

Total Kjeldahl nitrogen ••••••••••••••••••••••••••••• 54

Ammonia nitrogen •••••••••••••••••••••••••••••••••••• 55

Nitrite nitrogen •••••••••••••••••••••••••••••••••••• 55

Nitrate nitrogen •••••••••••••••••••••••••••••••••••• 55

Dissolved oxygen •••••••••••••••••••••••••••••••••••• 55

Oxygen uptake rate •••••••••••••••••••••••••••••••••• 55

Alkalinity •••••••••••••••••••••••••••••••••••••••••• 56

OXYGEN
TRANSFERSTUDIES••••••••••••••••••••••••••••••••••• 56

Laboratory Apparatus ••••••••••••••••••••••••••••••••••• 56

Nonsteady State Testing Procedure •••••••••••••••••••••• 57

Steady State Testing Procedure ••••••••••••••••••••••••• 58

IV. MATHEMATICAL
MODELING
•••••••••••••••••••••••••••••••••••••••• 60

BIOKINETICEQUATIONS
•••••••••••••••••••••••••••••••••••••• 61

STOICHIOMETRIC
EQUATIONS
•••••••••••••••••••••••••••••••••• 63

Qualitative Reactions •••••••••••••••••••••••••••••••••• 64

Quantitative Reactions ••••••••••••••••••••••••••••••••• 64

Carbon Limitations ••••••••••••••••••••••••••••••••••••• 66

Oxygen Limitations ••••••••••••••••••••••••••••••••••••• 68

Sl11-1l1.A.RY
••••••••••••••••••••••••••••••••••••••••••••••••• 69

v. RESULTS ••••••••••••••••••••••••••••••••••••••••••••••••• 71

EXPERIMENTAL
ANDTHEORETICAL
RESULTS
•••••••••••••••••••••• 78

Chemical Oxygen Demand••••••••••••••••••••••••••••••••• 78

vii
TABLEOF CONTENTS(cont'd.)

Page

Mixed Liquor Suspended Solids Concentration •••••••••••• 85

TKN and NH3-N Removal Efficiencies ••••••••••••••••••••• 85

Percent Distribution of Effluent Nitrogen •••••••••••••• 91

Destruction of Alkalinity •••••••••••••••••••••••••••••• 94

Biokinetic Constant Determination •••••••••••••••••••••• 96

OXYGENTRANSFERSTUDYRESULTS••••••••••••••••••••••••••••• 96

Oxygen Transfer Coefficients (KLa) ••••••••••••••••••••• 99

Oxygen Uptake Rates (R) •••••••••••••••••••••••••••••••• 99

Relationship Between Oxygen Uptake Rate and KLa••••••• 103

Alpha Coefficient (a) ···•••·•••••••••·••••••·•··•••••103


Beta Coefficient (S) ••••••••••••••••••••••••••••••••• 106

VI. DISCUSSION •••.•••••••••••••••••••••••••••••••••••••••••••. 107

DISCUSSIONOF EXPERIMENTAL
ANDTHEORETICALRESULTS••••••• 107

Chemical Oxygen Demand •••••••••••••••••••••••••••••••• 107

Mixed Liquor Suspended Solids Concentration ••••••••••• 109

TKN and NH3-N Removal Efficiencies •••••••••••••••••••• 110

Percent Distribution of Effluent Nitrogen ••••••••••••• 113

Destruction of Alkalinity ••••••••••••••••••••••••••••• 117

Biokinetic Constant Determiriation ••••••••••••••••••••• 118

DISCUSSIONOF OXYGENTRANSFERSTUDYRESULTS•••••••••••••• 120

Oxygen Transfer Coefficients (KLa) •••••••••••••••••••• 120

Oxygen Uptake Rates (R) ••••••••••••••••••••••••••••••• 122

viii
TABLEOF CONTENTS(cont'd.)

Relationship Between Oxygen Uptake Rate and KLa••••••• 123

Alpha Coefficient (a) ••••••••••••••••••••••••••••••••126


Beta Coefficient (S) ••••••••••••••••••••••••••••••••• 127

SUMMARY••••••.••••.••••••••••••••.•••••••••.•••••••••• • 128

VII. CONCLUSIONS ••••••••••••••••••••••••••••••••••••••••••••••••133

VIII. RECOMMENDATIONS
FOR FUTURESTUDY
•••••••••••••••••••••••••••• 135

IX. REFERENCES • •.••••••••••••••••••••••••.•.••..••••••••••••• • 136

APPENDIX A•••••••••••••••••••••••••••• •••••••••••••••••••••••••••• • 14 7,

APPENDIX B • ••••••••••••••••••••••••••••••••••••••••••••••••••••••• • 162

VITA ••••••••••••••••••••••••••••••••••••••••••••••••••••••••••••••• 193

ABSTRACT

:ix
LIST OF TABLES

Table

I Parameters Monitored Daily •••••••••••••••••••••••••••••••• 43

II Composition of Influent to Reactor 1 •••••••••••••••••••••• 47

III Composition of Influent to Reactor 2 •••••••••••••••••••••• 48

IV Biokinetic Coefficients Used in Mathematical Model •••••••• 74

V Mathematical Model Parameters ••••••••••••••••••••••••••••• 75

VI Summary of Steady State Data for COD:TKN = 6.07:1 ••••••••• 79

VII Theoretical Data Generated for COD:TKN = 6.07:1 ••••••••••• 81

VIII Summary of Steady State Data for COD:TKN = 0.65:1 ••••••••• 82

IX Theoretical Data Generated for COD:TKN = 0.65:1 ••••••••••• 84

X Summary of Oxygen Transfer Data for

COD:TK.N
= 6.07:1 ••••••••••••••••••••••••••••••••••••••• 100

XI Summary of Oxygen transfer Data for

COD:TKN= 0.65:1 ••••••••••••••••••••••••••••••••••••••• 101

X
LIST OF FIGURES

Figure

1 Flow Diagram for a Completely Mixed, Sludge Recycle


Activated Sludge Wastewater Treatment Process •••••••••••• 3

2 Bench Scale Activated Sludge Unit with Internal


Cell Recycle ••••••.••••••.•.••.•.•.•.••••••••••••••••••• 46

3-a Actual COD Removal Efficiency Versus Mean Cell


Residence Time •••••.•••.••••••.•.••••••••••••••.•••••••• 86

3-b Theoretical COD Removal Efficiency Versus Mean Cell


Residence Time ••••..••••.•••..••.••••••••••••••••••••••• 86

4-a Actual Soluble Effluent COD Concentration Versus Mean Cell


Residence Time ••••••••.•••••...•...•••.••••••••••••••••• 87

4-b Theoretical Soluble Effluent COD Concentration Versus


Mean Cell Residence Time ••••.•.•.•••••••.••••••••••••••• 87

5-a Actual Aeration Basin MLSS Concentration Versus Mean


Cell Residence Time ••••.••••..•••••••••••••••••••••••••• 88

5-b Theoretical Aeration Basin MLSS Concentration Versus


Mean Cell Residence Time ••••.•••.•.••••••••••••••••••••• 88

6 TKN Removal Efficiency Versus Mean Cell


R~sidence Time •.••••••••••.•..•....•..••••••••••••••••.. 89

7-a Actual NH3-N Removal Efficiency Versus Mean Cell


Residence Time ..•••••••••••••......••••••••••••••••••••• 90

7-b Theoretical NH3 -N Removal Efficiency Versus Mean Cell


Residence Time ••.••••.•.•••••••..•.••••••••••.•••••••••• 90

8-a Effect of Mean Cell Residence Time on Actual Percent


Distribution of Nitrogen for COD:TKN = 6.07:1 ••••••••••• 92

8-b Effect of Mean Cell Residence Time on Theoretical Percent


Distribution of Nitrogen for COD:TKN= 6.07:1 ••••••••••• 92

9-a Effect of Mean Cell Residence Time on Actual Percent


Distribution of Nitrogen for COD:TKN = 0.65:1 ••••••••••• 93

9-b Effect of Mean Cell Residence Time on Theoretical Percent


Distribution of Nitrogen for COD:TKN = 0.65:1 ••••••••••• 93

xi
LIST OF FIGURES (cont'd.)

Figure

10 Alkalinity Destroyed mg/1 Caco 3 /mg/l TKN Removed Versus


Mean Cell Residence Time •••••••••••••••••••••••••••••••• 95

11 Relationship of Alkalinity Predicted by Equation [4] to


Measured Alkalinity Change •••••••••••••••••••••••••••••• 97

12 Specific Growth Rate Versus Specific Substrate


Utilization Rate •••••••••••••••••••••••••••••••••••••••• 98

13 Effect of Mean Cell Residence Time and COD:TKNRatio on


Oxygen Uptake Rate ••••••••...••••••..•••••••••••••••••• 102

14-a Relationship Between Oxygen Uptake Rate and Oxygen


Transfer Coefficient •••••••••••••••••.•.••••••••••••••• 104

14-b Relationship Between Specific Oxygen Utilization Rate and


Oxygen Transfer Coefficient •••••••••••••••••••••••••••• 105

15 Oxygen Uptake Rate Versus KLa Corrected to


Standard Conditions ••.•••••.••••.•••••••••••••••••••••• 124

16 Oxygen Uptake Rate as a Function of KLa•••••••••••••••••• 125

xii
I. INTRODUCTION

The activated sludge process has a long and successful history in

the treatment of domestic and industrial wastewaters. Ardern and

Lockett [l] developed the activated sludge process in England in

1914. Since its conception, several modifications have been made to

the conventional process. Utilization of this process is widespread

although many of the mechanisms that make it work are still relatively

misunderstood. A resurgence of interest in biological treatment may

be attributed to the high cost associated with physico-chemical

schemes and an increased emphasis on the application of biological

sludges to farmlands for supplemental nutrients and enhancement as a

soil conditioner.

There are two major steps which characterize the activated sludge

process. The first involves substrate utilization and the second, a

solids liquid separation. Wastewater flows into the aeration basin

where it is brought into contact with a heterogeneous culture of

microorganisms which utilizes the organic constituents and other

nutrients for growth and energy requirements. As a result, substrate

is utilized and new bacteria are produced. Separation of the micro-

bial biomass from the liquid wastewater through secondary clarifica-

tion is accomplished in the second step. To achieve a required level

of treatment, it is necessary to operate the process at a specified

net growth rate (mean cell residence time). This can be accomplished

by wasting sludge from the secondary clarifier or directly from the

1
2

aeration basin. At steady state conditions, the amount of biomass

that is wasted from the system is equal to the quantity of biomass

that is synthesized in the system. A schematic diagram of the com-

pletely mixed activated sludge process is shown in Figure 1.

NEED FOR STUDY

Mathematical descriptions of the activated sludge process along

with stoichiometric relationships have been presented to help promote

a unified approach to describing the process [2-10]. Use of these

models has made a significant contribution to our understanding of

some of the principles and mechanisms involved in process design and

operation.

With the increased emphasis on biological treatment and the acti-

vated sludge process in particular, there is a need to evaluate and

understand some of the most basic mechanisms that affect the micro-

organisms present. For efficient biological oxidation to take place,

the microbial biomass in the aeration basin must be provided with

sufficient quantities of nutrients and oxygen so these will not be

growth limiting. Oxygen transfer is one area which requires further

investigation. In particular, the problems encountered in the process

of selecting, specifying, and testing aeration equipment needs to be

addressed. Also, operation of the activated sludge process can result

in a situation where dual substrate limitations may exist. It has

been shown theoretically [11-13] that different substrates may control

the growth rate at various mean cell residence times. At low mean
AERATION BASIN SECONDARY CLARIFIER

INFLUENT EFFLUENT

V,X

SLUDGE RETURN WASTE SLUDGE


Qw,Xr

Figure 1. Flow Diagram for a Completely Mixed, Sludge Recycle Activated Sludge Wastewater
Treatment Process.
4

cell residence times, a particular substrate will restrict growth

while at longer residence times a different substrate may be growth

limiting.

OBJECTIVES

The specific objectives of this research investigation were to:

1) Investigate the operation of the completely mixed activated

sludge process as it approached a dual substrate limitation

with carbon being the limiting nutrient at low mean cell

residence times and dissolved oxygen being growth limiting

at higher mean cell residence times.

2) Examine the use of biokinetic and stoichiometric equations

for predicting effluent quality when operation of the acti-

vated sludge process approaches a dual substrate limita-

tion.

3) Determine and compare the oxygen transfer coefficient (KLa)

obtained from steady state and no~steady state aerator test-

ing procedures.

4) Elucidate the relationship between· the oxygen uptake rate of

the mixed liquor suspended solids and the oxygen transfer

coefficient.
II. LITERATUREREVIEW

Utilization of the activated sludge process for the treatment of

domestic and industrial wastewaters has been widely practiced. Unfor-

tunately, several of the mechanisms that affect the efficiency of the

process are still not clearly understood. Successful operation of the

process requires an understanding of three important factors. First,

a knowledge of growth kinetics under dual substrate limitation where

organic carbon controls the growth at a specified mean cell residence

time and dissolved oxygen at another is required. The nitrification

process and the parameters that affect it must be understood if a

nitrified effluent is desired. A third consideration is the transfer

of oxygen to the process which must meet total oxygen requirements and

provide adequate mixing when diffused aeration is utilized. In this

chapter a thorough review of the literature will be presented for

technical background with regards to dual substrate limitations,

nitrification, and oxygen transfer.

CONVENTIONAL
BIOKINETICTHEORY

Most of the conventional biokinetic theories that have been

developed to mathematically model the completely mixed activated

sludge process have been used with success. These mathematical des-

criptions of the processes used in conjunction with stoichiometric

relationships have been presented to help advance the design and oper-

ation of the process [2-10]. Of course there are several inherent

5
6

problems associated with these models including the following assump-

tions or limitations: 1) steady state conditions exist, 2) influent

wastewater characteristics remain constant, 3) effluent substrate

concentration is independent of influent substrate, 4) no accumulation

of sludge in secondary clarifiers, 5) negligible concentration of

microorganisms entering the aeration basin, and 6) only soluble efflu-

ent substrate concentration can be predicted. These restrictions are

not severe if the model is used in the proper perspective. Activated

sludge treatment plants fail to meet NPDES (National Pollution

Discharge Elimination System) permit regulations and/or stream stan-

dards occasionally due to solids carry over in the secondary efflu-

ent. Part of the problem lies in state and federal bureaucracies

which have set unrealistic effluent standards without an understanding

of the purpose of biological treatment plants which is to reduce the

readily biodegradable organics in the wastewater to low levels [14].

The conventional biokinetic theory for the completely mixed acti-

vated sludge process is based on single nutrient substrate limita-

tions. Organic carbon usually serves as the growth limiting substrate

while all other nutrients and trace elements are considered to be

provided for in excess. Most of the studies involving substrate limi-

tations in biological systems have been conducted in continuous flow

chemostats with pure microbial cultures and involved only single

nutrient restrictions [15-17]. These investigations have established

the degree to which the composition of a microbial cell can be con-

trolled under a single nutrient limitation. Chemostat studies


7

utilizing pure cultures have provided considerable data on single

substrate limitations but their value is somewhat limited since full

scale plants usually employ recycle and heterogeneous cultures. Since

single essential nutrients were found to affect microbial growth in

different ways, investigators were prompted to elucidate the relation-

ships between cell growth and multiple nutrient or substrate limita-

tions [17].

DUALSUBSTRATELIMITATIONS

It is possible for dual substrate limitations to occur in biolog-

ical systems. Instead of one substrate controlling the growth rate

and/or the extent of growth of the biomass, two substrates affect the

net growth rate of the microorganisms. Two schools of thought exist

concerning multiple substrate limitations: an interactive model or a

non-interactive model. Both are derived from the single substrate

limiting or conventional biokinetic models and each can be used to

define the conditions where dual substrate limitations may occur. The

interactive model assumes that when two substrates are present at

concentrations below their saturated values then both must affect the

overall growth of the microorganisms. This school of thought repre-

sents a "simultaneous" substrate limitation at a given time. The non-

interactive model is differentiated from the interactive model in that

the rate of growth of the microorganisms can only be limited by one

particular substrate at a givell tJme.


8

Models for Dual Substrate Limitations

Several models have been developed to show the relationship

between the growth rate of the microorganism and the limiting sub-

strate. Sykes [18] presented a model based on the Monod equation. It

was reported that a given growth medium may contain more than one

limiting nutrient and that different nutrients may be limiting at

different stages of the growth cycle.

A comparison of the single limiting substrate model and the mul-

tiple limiting substrate model was made by Sykes [19]. For situations

involving multiple substrate limitations in a chemostat, it was shown

that increasing the influent concentration of a given nutrient will

increase its ambient concentration, while decreasing the ambient con-

centration of all other nutrients. The biomass concentration will

also increase. Sykes also reported that batch cultures are incapable

of discerning multiple nutrient limitations.

Ryder and Sinclair [20] developed a simple model for the contin-

uous culture of microorganisms in fermenters under both oxygen and

feed substrate limitations. Their non-interactive model lists a set

of equations for feed substrate limitations and another set of equa-

tions for oxygen limiting conditions. Experimental results where

Candida utilis was grown on glucose verified their theory. The model,

however, suffers from a failure to account for endogenous metabolism

and cell death.

A comparison of the interactive model proposed by Frederickson et

al. [21] and the non-interactive model proposed by Ryder and


9

Sinclair [20] was made by Sinclair and Ryder [22] for growth of

Candida utilis utilizing glycerol as substrate. An endogenous

metabolism term after Herbert [23] was incorporated into each model

for the comparison. Both models were found to be satisfactory in

predicting the culture of microorganisms under both carbon and oxygen

limiting conditions.

The model proposed by Ryder and Sinclair [20] was reviewed by

Bader et al. [24]. The basic criticism of the model was that dual

substrate limitations were not assumed to be simultaneous or inter-

active at a given time and that the model would probably work well in

chemostats but in other situations would fail. Bader et al. [24]

promote the interactive model proposed by Megee et ~- [25].

Chemostat studies which examined the growth of Enterobacter

aerogenes under single and dual substrate limitations were conducted

by Cooney~~- [17]. Carbon to nitrogen and phosphate to nitrogen

ratios of the influent as well as dilution rate were varied to inves-

tigate these effects on changes in the macromolecular cell composi-

. tion. Acetic acid and metabolite production was found to be dependent

on both the nutrient limitation and dilution rate.

The transient response of a culture of Enterobacter aerogenes

following a pulse injection of ammonia to a chemostat operating under

a dual substrate limitation of nitrogen and phosphorus was conducted

by Cooney and Wang [26]. These investigators were able to show that

the Monad model does not describe growth under transient conditions.

Also, microbial cells operating under a double substrate limitation


10

may have a reserve biosynthetic capacity which allows immediate

adjustments to small changes in a pulse or shock loading. Increasing

the phosphate limitation decreased the cells ability to respond to a

release from the nitrogen restriction.

Recently, a comparison of the interactive and non-interactive

models was made by Bader et al. [27]. The authors first discuss the

basic single substrate limiting models based on the Monod, Blackman,

and exponential kinetic models. These were incorporated into the two

dual substrate limiting models for comparison. They also allude to

the concept that one substrate may control the rate of growth and

another may control the extent of growth. An enzyme analog was pre-

sented to show that both models could be utilized under specific con-

ditions. The final conclusion was that a simple and unique kinetic

model which handles all types of dual substrate limitations does not

exist.

Dual Substrate Limitations in Biological Treatment

Most of the research conducted on the activated sludge process

has focused on single substrate or single nutrient limitations.

Removal of nutrients from wastewater, primarily nitrogen and phos-

phorus, has received the greatest attention since they stimulate cul-

tural eutrophication of natural bodies of water. The interaction of

limiting nutrients in a heterogeneous culture such as activated sludge

is a relatively unexplored field.


11

There are reports in the literature where nutrient interactions

have been considered. Several studies were referenced by Sykes [18,

19] where nutrient interrelationships were accounted for in algal

cultures. For most wastewaters carbon is the limiting nutrient and

phos~horus is present in excess [28, 29]. The investigation of phos-

phorus removal in the activated sludge process by Stall and Sherrard

[7] promoted the idea of dual substrate limitations existing in the

activated sludge process. These investigators found that phosphorus

could become a limiting nutrient at high COD:P ratios and low mean

cell residence times (eC ). At high eC values, the organic carbon as

measured by CODwould be the limiting substrate.

Grady and Lim [30] give a brief description of the role of dual

substrate limitations in the denitrification process. A double Monod-

type function was presented and can be used successfully to model dual

substrate limitations. They aiso sugg~st that biological systems

limited by oxygen and organic carbon could be modeled with a double

Monod-type function.

Recently, several papers have been presented which discuss non-

interactive dual substrate limitations in the activated sludge pro-

cess. Theoretical examples were used to illustrate how standard bio-

kinetic theory and stoichiometry can be employed to determine where

dual substrate limitations may occur. Hawkins and Sherrard [11] have

shown theoretically that phosphorus can be the limiting nutrient at

high COD:P ratios and low mean cell residence times. Nitrogen limit-

ing conditions will exist in the activated sludge:process when


12

+-N ratios
high COD:NH and low mean cell residence times are utilized
4
[12]. When the oxygen requirements of the activated sludge exceed the

oxygen transfer capability of the aeration equipment the system will

be operating under dual substrate limitations. Mines and Sherrard

[13] have described a biokinetic stoichiometric model which can be

used to predict the effluent quality from an activated sludge plant

operating under an oxygen restraint. At low mean cell residence times

carbon is the limiting nutrient and at high 0 C values oxygen will be

nutrient limiting. Each of the above studies leads to the conclusion

that dual substrate limitations are dependent on the mean cell resi-

dence time (reciprocal of the net microorganisms specific growth rate)

and the ratio between the two limiting nutrients involved.

Stenstrom and Poduska [31] have attributed the broad range of

dissolved oxygen concentrations required for nitrification found in

the literature to simultaneous duai substrate limitations. A double-

Monod type equation was utilized to illustrate the mean cell residence

times where both ammonia and dissolved oxygen can be rate limiting.

Importance of Nitrification

In combined carbon/nitrogen oxidation systems such as the acti-

vated sludge process, nitrification of the wastewater can be accom-

plished if the proper environmental conditions are maintained.

Several excellent reviews on nitrification can be found in the litera-

ture but only the most important parameters that affect nitrification

pertinent to this research will be presented here [32-35].


13

Nitrification involves a group of chemoautrophic bacteria, collec-

tively known as the nitrifiers, which oxidize ammonium nitro-

gen (NH!-N) to nitrate nitrogen (No;-N) in a two-step sequential

reaction. Heterotrophic nitrification has also been shown to occur

but is only significant in environments that have very alkaline or

acidic pH values [33].

The first reaction involves the oxidation of NHt-N to nitrite

nitrogen (No;-N) by the genus Nitrosomonas. Focht and Chang [33]

state that oxygen is utilized in this reaction for two purposes: 1)

one oxygen atom is directly incorporated into the substrate and, 2)

oxygen serves as an electron acceptor during electron transfer through

the cytochrome system. There are several intermediate nitrogenous

compounds that have been proposed to exist between ammonium and

nitrite, the most noteworthy, hydroxylamine and nitroxyl [33-34, 16-


37). Much confusion exists in the literature as to which inter-

mediates actually occur and whether or not the reactions are chemical

or biological in nature. The following microorganisms are also cap-

able of oxidizing ammonium to nitrite: Nitrosococcus, Nitrosospira,

Nitrosolobus, Nitrosocystis, Nitrosogloea, and Nitrosospima but are of

limited importance in biological oxidation systems [34, 38).

In the second step of the sequential reaction, Nitrobacter oxi-

dizes N02 -N to N03 -N. This step is simpler and better understood.

Oxygen serves as a terminal electron acceptor and the oxygen atom

incorporated into nitrate is generated from water and not from oxygen
14

[33,37]. Two other genera of nitrite oxidizers include Nitrospina and

Nitrococcus which are of secondary importance [34].

Both reactions are thermodynamically favorable with estimates of

free energy of release of between 58 and 84 kilocalories per mole for

the ammonium oxidation and between 15.4 and 20.9 kilocalories per mole

for nitrite oxidation [36, 39]. The individual steps and the overall

equation are summarized by the following reactions [36].

Nitrosomonas >2 NO-+ 4H+ + 2H 0 [l]


2 2

Nitrobacter > 2 NO [2]


3

Nitrifiers > NO + 2H+ +HO [3]


3 2

Rarely has a buildup or accumulation of nitrites occurred in aerobic

treatment systems since the energy yield of the equation [l] is much

greater than that of equation [2]. Therefore, the nitrification rate

is dependent upon the ammonimn oxidation reaction [36, 40].

For effective nitrification to occur, a large population of

nitrifiers must be established in the aeration basin. Operation of

the activated sludge process at mean cell residence times greater than

the reciprocal of the maximmn growth rate of the nitrifiers will

ensure nitrification if the proper environmental conditions exist.

The nitrifiers grow slowly and generally require a sludge age greater

than two to three days for nitrification to occur. Mean cell


15

res~dence times less than these values normally result in the nitri-

fiers being "washed-out" of the system.

Environmental Conditions Needed. The nitrifying organisms grow

considerably slower than their competitors, the heterotrophs. In

wastewater treatment processes, there are several factors that influ-

ence the growth rate of the nitrifiers. Parameters affecting nitrifi-

cation rates include: oxygen requirements, temperature, pH, alka-

linity, and substrate and product inhibition.

Oxygen requirements. Dissolved oxygen concentration of the mixed

liquor is an important parameter that must be controlled if effective

treatment is to be accomplished. Standard theory states that the cell

respiration rate will proceed at a rate independent of the dissolved

oxygen concentration as long as the DOlevel remains above a specified

critical dissolved oxygen concentration [36, 42]. Wuhrman [41] pro-

posed a critical oxygen concentration of 0.1 mg/1. Various dissolved

oxygen levels have been reported in the literature which must be main-

tained to ensure nitrification [42]. The range starts at 0.1 mg/1 and

goes up to 3.0 mg/1 dissolved oxygen. After acclimation, nitrifi-

cation was not inhibited by high DO levels of 38 mg/1 using pure oxy-

gen [43]. Other investigators [40] recommend that nitrification

systems be designed at a DO level of 2.5 to 3.0 mg/1 under winter

conditions. Then the lower DO levels prevalent in summer conditions

may be offset by the increased nitrifier activity at the warmer


16

temperatures. According to Benefield and Randall [36] the dissolved

oxygen concentration for a combined carbon oxidation/nitrification

process would require a DO of 2.0 mg/1 and a separate stage system

would only require a DO level greater than 1.0 mg/1. The consensus of

most researchers would be that 2.0 mg/1 of dissolved oxygen should be

maintained to ensure successful carbon oxidation/nitrification.

From stoichiometric considerations, Equations [1] and [2] require

4.57 mg of oxygen per milligram of NH:-N oxidized for complete nitri-

fication. Other researchers have found that the actual ratio of oxy-

gen utilized to nitrogen oxidized to be somewhat less than that pre-

dicted stoichiometrically [39, 44-45]. Sharma and Ahlert [37], citing

Kiesow [46] stated that variable stoichiometry exists for nitrite

oxidation (nitratification) which explains the variable amounts of

oxygen being reported for utilization. The amount of ammonia that can

be removed will be directly proportional to the amount of oxygen that

is supplied to the system in excess of the carbonaceous oxygen demand

[47]. Oxygen requirements for substrate biooxidation will be deter-

mined by the nitrogen concentration in relation to the carbon source

and is inversely affected by the mean cell residence time [48].

According to Sherrard [49], total oxygen requirements for activated

sludge processes are not a constant value and are dependent on the

COD:TKNratio of the wastewater and mean cell residence time.

Another important consideration is that autotrophs are out-

competed for oxygen by the heterotrophs [34]. Between Nitrosomonas

and Nitrobacter, Belser [35] citing Laudelout al. [50] stated that
17

for all practical purposes, the ammonium oxidizers out-compete the

nitrite oxidizers to the point of completely inhibiting them. This

observation suggests that if aeration capacity is exceeded then

nitrite accumulation could result.

Temperature. Several studies [33, 34, 37) have reported nitrifi-

cation to occur over the temperature range of 4 to 42°c. Optimal

temperature for nitrification ranges between 28 and 36°c [33, 34,

37). Nitrification kinetics usually conform to the Arrhenius equation

[33, 37). Various equations [36, 39) have been proposed to relate

growth rate to temperature. Lower temperatures seem to have a more

drastic effect on nitrifier growth than do higher temperatures. In

the temperature range from 5 to 35°c, the rate of nitrification

increased as temperature increased [51). If temperatures less than 12

to 1s0 c are likely to be experienced in the aeration basin, Adams and

Eckenfelder [40) suggest that biological nitrification may not be

feasible.

pH and alkalinity. Both Nitrosomonas and Nitrobacter are

affected by the pH of the mixed liquor. Nitrification will generally

proceed over the pH range of 6.0 to 9.0 [38). Optimum pH occurs at a

neutral to slightly alkaline pH (7.0 - 8.0) [33, 34). Both the free

ammonia concentration and undissociated nitrous acid concentration

have been found to inhibit nitrification and are pH dependent


18

[52-55]. Substrate and product inhibition will be discussed in more

detail in a subsequent section.

As nitrification proceeds, hydrogen ions are released in the

mixed liquor as NH+


4 ~N is oxidized. Autotrophic bacteria utilize inor-

ganic carbon for cellular synthesis. Sufficient alkalinity must be

present in the wastewater to provide the buffering capacity necessary

to prevent a shift in pH and serve as the inorganic carbon source. If

insufficient alkalinity is present, nitrification will be inhibited.

Sherrard et al. [56] have proposed and verified experimentally the

following equation which can be used to predict the amount of alka-

linity destroyed or produced in an aerobic treatment system:

bAlk = 3.57 [(bfiltrate organic-N) - (synthesized-N)]

Substrate and product inhibition. Some organic compounds such as

thiourea, halogenated solvents and phenolic compounds, phenol,

cyanides and cresol are potentially toxic to activated sludge [57].


Keenan et al. citing reference [39], state that high substrate concen-

trations (COD= 939 mg/1 and BOD·= 118 mg/1) in the mixed liquor may

have inhibited nitrification [65]. Low organic loadings were used to

maintain mean cell residence times that were favorable to the slow-

growing nitrifiers [58]. Under normal circumstances, domestic waste-

water does not contain high organic or nitrogenous concentrations that

would inhibit nitrification~ Industrial, poultry, and agricultural


19

wastewaters may contain high concentrations of each that would be

inhibitory.

Both Nitrosomonas and Nitrobacter are sensitive to their own

substrate but even more so to the substrate of the other [37, 54].

Anthonisen et al. [54] quoting Meiklejohn [59] stated that the respir-

ation and growth of Nitrosomonas is depressed by nitrite and that

Nitrobacter is sensitive to the ammonium ion but even more so to free

ammonia. Nitrate was found to be only slightly toxic to both

Nitrosomonas and Nitrobacter.

Several investigators [50, 52-55, 60-64] have reported on treat-

ment systems where incomplete nitrification and a large accumulation

of nitrites occurred. Inhibition of nitrification was found to be

dependent on the concentration of free ammonia (FA) and free nitrous

acid (FNA) present. Free ammonia concentrations in the range of 10 to

150 mg/1 were found to be inhibitory to Nitrosomonas and FA concen-

trations of 0.1 to 1.0 mg/1 were inhibitory to Nitrobacter, while FNA

in the range of 0.22 to 2.8 mg/1 inhibited both [54]. Operating data

from a full scale activated sludge plant treating high nitrogenous

wastes (286 mg/1 ammonia nitrogen) indicated that substrate inhibition

due to ammonium ion, however, product inhibition due to nitrite and

nitrate was not considered [65]. A total concentration of NR3

and NH:-N of 60 mg/1 at a pH of 8.5 did not inhibit nitrification in

studies conducted by Wild et al. [66]. Kholdebarin and Oertli [67]

presented results indicating that nitrite oxidation is stimulated by

the addition of NH+


4-N. At the nitrogen levels employed in their study
20

(2.7 and 28.0 mg/1 N) nitrite oxidation could be stimulated but at

much higher levels, substrate inhibition would probably result as can

be inferred from the study by Keenan et al. [65]. Charley~ al. [43]

found that nitrite concentrations greater than 20 mg/1 were inhibitory

to nitrification. Nitrobacter was found to be sensitive to ammonia

concentrations greater than 1.0 mg/1 as N and to nitrous acid concen-

trations greater than 2.8 mg/1 as N [55]. In their study, Boon and

Laudelout [60] indicated that high concentrations of nitrate noncompe-

titively inhibit Nitrobacter. A N02 -N concentration of 1400 mg/1 was

found to cause a 40% inhibition in Nitrobacter activity.

Kinetics and Biokinetic Constants. Several excellent reviews on

nitrification kinetics can be found in the literature [33-35, 37, 43,

57, 68, 69]. Microbial nitrification kinetics have been found to fit

the Michaelis-Menten Model [43, 71]. Other researchers [51, 70] have

found that both ammonium oxidation and nitrite oxidation kinetics fol-

low zero order reactions. The Michaelis-Menten equation reduces to

zero order at high substrate concentrations when the substrate concen-

tration is much greater than the saturation constant Ks• The studies

conducted by Wong-Chong and Loehr [70] and Srinath et al. [51] were

conducted at substrate concentrations ranging from 100-1000 mg/1 N and

50-1500 mg/1 N which should follow zero order kinetics. Wu [48] con-

cluded from his studies that substrate utilization rate (q), yield

coefficient (Y), and endogenous decay coefficient (kd) are not


21

affected by the concentration of nitrogen in the feed in completely

mixed systems with cellular recycle.

The saturation constant for the Michaelis-Menten model was

reported to range from 1 to 10 mg/1 N for ammonia oxidation and 5 to 9

mg/1 N for nitrite oxidation [33, 34). For full scale activated

sludge plants, Poduska and Andrews [68) report that saturation con-

stants for Nitrosomonas and Nitrobacter are approximately 1 mg/1 N.

Lawrence and McCarty [2] list Ks values ranging from 0.6 to 3.6 mg/1 N

for ammonium oxidation and 0.3 to 4.7 mg/1 N for nitrite oxidation.

Cell yield values for Nitrosomonas and Nitrobacter have been

reported to range from 0.03-0.13 and 0.02-0.08 weight of cells per

weight of energy [37). Growth yield coefficients in mg biomass per mg

N were determined to range from 0.05-0.29 and 0.02-0.084 for

Nitrosomonas and Nitrobacter [2].

The energy maintenance coefficient (kd) was reported to be

0.05 days- 1 for both species [2]. Values listed for the maximum sub-

strate utilization per unit weight of microorganisms (k) ranged from

0.9 to 30 and 3.9 to 100 mg N/mg-day for ammonium and nitrite oxida-

tion respectively [2].

OXYGENTRANSFER

The process of selecting, specifying, and testing aeration equip-

ment has generated considerable interest in recent years. With more

emphasis on energy conservation, engineers must develop efficient and

economical means of transferring oxygen into biological oxidation


22

systems. A review of the literature shows that there are several

misunderstandings of the basic fundamentals regarding the rates and

mechanisms involved in oxygen transfer in wastewater systems. Numer-

ous articles state that most of the aeration systems in operation

today are either overdesigned or underdesigned [72-74].

Many biological reactors are operated in such a fashion as to

provide "an excess amount of air" to provide a dissolved oxygen con-

centration greater than a stated minimum value. In most cases this is

wasteful and the process may in fact operate substantially more effi-

ciently on a energy basis at a lower DO level. The dissolved oxygen

concentration in the aeration basin of an activated sludge unit is of

prime importance. Oxygen is an essential nutrient required by aerobic

microorganisms indigenous to the activated sludge process. If biolog-

ical treatment processes are to accomplish their objectives effec-

tively, an adequate air supply must be maintained to satisfy: oxygen

demand of the waste, oxygen requirements for incorporation into cellu-

lar material, endogenous respiration of the microbial biomass and suf-

ficient mixing to keep the sludge in suspension.

The lack of a consensus standard for the testing of aeration

equipment has also contributed to a renewed interest in oxygen trans-

fer. Currently, the American Society of Civil Engineers (A.S.C.E.)

Subcommittee on Oxygen Transfer Standards is working on an interim

transfer standard. A special workshop sponsored by the U. S.

Environmental Protection Agency and the A.S.C.E. was held to summarize

the existing and historical development in the testing and evaluation


23

of aeration equipment [75]. Unlike some manufactured equipment

(blowers and pumps), aerators are not tested aga;i.~.s~, a _c;onsensus

standard. There is a need for a consensus standard and a review of

oxygen transfer theory and testing procedures.

Theory

The most widely used and accepted theory which describes mass

transfer was proposed by Whitman [76]. Two other more comple~ models

have been proposed which describe oxygen transfer and involve surface

renewal parameters. Higbie [77] modified the Whitman model to account

for a time-dependent concentration profile in the liquid film

(Penetration Theory). Higbie's model was further refined by

Danckwerts [78] to allow for a statistical variation of the exposure

of fluid elements to gas. Utilization of these models has not gained

wide acceptance due to their complexities, even though the equations

developed may describe the transfer of oxygen from air to water more

realistically.

Since the two-film theory of gas-liquid transfer has gained wide

acceptance in the environmental engineering community, .it will be

emphasized in this study. The two-film theory involves physical mass

transport of a gas across a two-film layer consisting of a gas film

and a liquid film. Fick's First Law of Diffusion can be used to model

molecular diffusion as follows:

ilM
-= - [5]
at
24

where: :: = mass transfer rate, mass/time,

DL = diffusivity constant, area/time,

A = cross-sectional area across which the solute is

diffusing, area, and

-ac = concentration gradient, mass/volume-time.


ax

When steady state conditions exist, the following two Equations

([6] and [7]) can be used to determine the transport rate across the

gas and liquid films [79].

D A
V p
[6]
RTs-
pb

[7]

dn
where: dt a = rate of moles of gas A diffusing from point 1 to

point 2, lb-mole/hr,

Dv = molecular diffusivity in gas, ft 2 /hr,

n1 = molecular diffusivity in liquid, tt 2/hr,


A= cross-sectional area which diffusion is taking

place, ft 2 ,

p = total pressure, atm,

pal= partial pressure of diffusing gas at point 1,

atm,
25

pa 2 = partial pres~ure of diffusing gas at point 2,

atm,

pb = partial pressure of nondiffusing gas, atm,

T = absolute temperature, 0 R,

R= universal gas constant, 0.729 ft 3-atm/lb mole- 0 R,

S = distance between points 1 and 2, ft,

cal= diffusing molecular species concentration at

point 1, lb-mole/ft 3,

ca 2 = diffusing molecular species concentration at

point 2, lb-mole/ft 3,

ca+b = total number of moles of species A and B per

unit volume, lb-mole/ft 3 , and

cb = concentration of diffusing molecular species B, lb-

mole/ft3.

Molecular diffusion of a gas across the films and into the bulk liquid

can be expressed as follows:

[8]

where: kG = mass transfer coefficient for gas, lb-mole/hr-ft 2-

atm,

kL = mass transfer coefficient for liquid, ft/hr,

pg = partial n.ressure of gas in-bulk gas phase, atm,

pi = partial pressure of gas at the interface, atm,

Ci = concentration of gas at the interface, lb-

mole/ft 3 , and
26

C = concentration of gas in bulk liquid, lb-mole/ft 3


.
For low soluble gases such as oxygen, the gas diffuses so slowly

through the liquid film that only a small concentration gradient is

required across the gas film. Therefore, the liquid film at the

interface becomes saturated with the gas at pressure Pg, and an equi-

librium condition exists between Ci and Pg. Pi is approximately equal

to Pg since a small concentration gradient exists across the gas

film. The DO concentration at the interface, Ci, actually represents

Cs, the saturation concentration of gas (oxygen) in the liquid film.

Since the liquid film controls the rate of diffusion, the gas film can

be neglected. The equation for physical mass transfer can be

rewritten as Equation [9] when the gas film is neglected.

[9]

Knowing that dM V dC
dt = dt

where: V = volume of the liquid phase, volume.

Substituting Equation [10] into Equation [9] results in the fol-

lowing expression:

[ 11]

Dividing both sides of Equation [11] by the volume of the aeration

basin yields the following equation:

[ 12]
27

Substituting Equation [13] into Equation [12] will result in Equation

[14] which represents the overall equation for physical mass transfer.

[ 13]

dC
-= [14]
dt

Aerator Test Methods

Several methods have been employed to measure the oxygen transfer

rate of aerators. Aeration equipment can be tested in clean tap water

or dirty water (wastewater) under nonsteady state or steady state

conditions. The most widely used nonsteady state methods involve

aeration of tap water, sulfite oxidation, and aeration of mixed liquor

or supernatant. Aeration of tap water, sulfite oxidation, and aera-

tion of mixed liquor, effluent, or supernatant have also been prac-

ticed under steady state conditions. Two lesser utilized techniques

include off-gas analysis and radioactive dye tracer studies. Descrip-

tions of these procedures can be found in the literature [42, 80-

83].

Usually, the nonsteady state reaeration test in clean tap water

is the method of choice, especially by aeration manufacturers. This

method is preferred since it is relatively simplistic in nature, low

in cost to perform, and tends to yield higher transfer rates than


28

other procedures employed. A brief summary of the major steps

involved in this testing procedure follows: 1) deoxygenation of a

given volume of water by bubbling nitrogen gas through it or addition

of sodium sulfite along with a catalyst such as cobalt chloride or

copper to enhance the deoxygenation reaction, 2) reaeration of the

sample to saturation, 3) record of DO concentration versus time, and

4) manipulation of the data graphically to determine the oxygen trans-

fer coefficient KLa.

Steady state aeration of tap water has been practiced but not to

a large degree. Difficulties encountered in maintaining constant flow

through rates and the quantity of water required have limited its

use.

Sulfite oxidation under steady and nonsteady state conditions has

been utilized historically in oxygen transfer studies. Major factors

limiting the use of this method involve chemical costs and the volume

of water required for steady state testing.

Dirty water testing involving the aeration of mixed liquor, mixed

liquor supernatant, raw wastewater and effluent wastewater has also

been utilized in KLa determinations under both steady and nonsteady

state conditions. Utilization of wastewater in oxygen transfer stu-

dies has several disadvantages (which will be discussed in a subse-

quent section) but more realistically relates to oxygen transfer in

the activated sludge process. Both the nonsteady and steady state

methods have been used but not as frequently as the clean water non-

steady state testing procedure.


29

Off-gas analysis and radioactive dye tracer techniques are two

relatively new methods for determining the KLa value. Both methods

are somewhat sophisticated and require special equipment if they are

to be employed in testing.

Several methods are available to select from for testing an aera-

tion device. There is still no firm evidence as to which of the above

methods yields the best results however, some of the procedures are

more realistic and simple to conduct, therefore they are more widely

utilized. The nonsteady state reaeration procedure in clean water has

been recommended to the A.S.C.E. Oxygen Transfer Standards

Subcommittee for incorporation into the proposed consensus standard

for aerator testing in clean water.

Alpha, Beta, and Theta

Aerators are normally tested by the manufacturer in field tests

conducted at standard conditions. Standard conditions for most manu-

facturers include the following environmental conditions: 1) clean

tap water, 2) zero dissolved oxygen, 3) one atmosphere of pressure,

and 4) 20°c. These parameters utilized in conjunction with the non-

steady state reaeration test result in a convenient testing procedure

that yields a higher oxygen transfer coefficient since there is a

larger driving force when the initial DO level is zero. Three param-

eters required by the design engineer to translate the data supplied

by the manufacturer to process conditions include alpha (a),

beta (6), and theta (e).


30

The alpha factor is utilized to relate the overall oxygen trans-

fer coefficient at standard conditions to that of process condi-

tions. This relationship can be expressed as follows:

~a wastewater
a = [15]
~a tap water

where: KLa = oxygen transfer coefficient, time- 1 •

To determine a, the aerator must be tested by one of the pre-

viously discussed techniques to determine the transfer rate in both

tap water and wastewater.

It is also necessary to correct the saturated DO concentration at

standard conditions to those at process conditions. The beta factor

is used to correct for dissolved solids and waste constituents in the

wastewater according to the following relationship:

DOsat wastewater
B= [16]
DOsat tap water

Beta is usually determined by taking identical volumes of tap water

and the actual wastewater and then aerating them until a constant

dissolved oxygen concentration (DOsat) is obtained. The ratio between

the two is the B factor or coefficient.

Temperature corrections from standard conditions to process con-

ditions can be accomplished by utilizing the theta factor (e). Nor-

mally, theta is used to correct for the appropriate temperature

according to the following equation:

[17]
31

where: T = temperature at process conditions, 0 c.


A wide range of e values have been reported ranging from 1.008 to

1.047 and a value of 1.024 is commonly used [84, 85].

Factors Affecting Transfer

The mass transfer of a gas through a gas/liquid film into the

bulk liquid has previously been described by Equation [14]. It must

be remembered that mass transfer is a physical or absorptive phenom-

enon. Some of the testing procedures for aerators involve chemical or

biological reactions. Therefore, Equation [14] must be modified or

other equations used. Major factors affecting oxygen transfer are

testing procedures, scaleup, a coefficient, e coefficient, and manip-

ulation of data. Parameters that influence the above variables

include turbulence and mixing intensity, mixed liquor suspended solids

concentration, surfactants, temperature, geometry of tank, and the

type of aeratiQn device.

Testing Procedure Difficulties. Depending on the testing proce-

dure used, the rate of a chemical or biochemical reaction may be

determined rather than the physical absorption capacity of the

liquid. Tsao [86] states that oxygen transfer coefficients are larger

when there is a simultaneous biological oxidation than when there is

not. As stated previously, the nonsteady state reaeration test in tap

water is the most widely used aerator testing procedure. Some


32

researchers [87] state that this does not truly represent a physical

process but involves a reaction between oxygen and the cobaltous and

sulfite ion. Busch [88] reported that the transfer potential of an

aerator tested at standard conditions of zero dissolved oxygen has

little to do with efficiency of field conditions unless the reaction

uses the full capacity of the aerator.

Oxygen transfer data obtained from sulfite oxidation have been

misinterpreted [42]. The oxygen transfer rates measured by the sul-

fite method will yield greater values than what would be measured in

pure water because of the chemical reaction involved. Nonsteady state

reaeration tests conducted in mixed liquor also present a problem.

Casey and Karmo [89] found that mixed liquor suspended solids did

influence the oxygen transfer in aeration systems. Two flocculent

sludges were analyzed and found to enhance the rate of oxygen transfer

to the same extent up to a concentration of 2000 mg/1. Above this

level, the a factor for their sludge A decreased whereas the a factor

for their sludge B increased with increasing solids concentration.

Other researchers [72] state that biological solids have a detrimental

effect on the rate of oxygen transfer.

Steady state aeration tests conducted in tap water or wastewater

are seldom used because~of the difficulty of maintaining steady state

conditions. In full scale testing, this is quite a difficult under-

taking because of the large volume of sample required and the con-

struction of testing facilities. McWhirter [90] states that steady

state testing of t~p water indicates that results are 10% higher than
33

those obtained from nonsteady state tests. Steady state problems in

activated sludge are numerous. If diffused aeration is being tested,

then other mechanical devices must be used to keep the solids in sus-

pension. Constant flowrate through the testing basin must be main-

tained and the wastewater characteristics must be the same. These

parameters are almost impossible to control if the aerator is tested

at the actual treatment plant.

Scaleup. If bench scale models are utilized in determining

transfer rates and a factors, caution must be exercised in applying

these data to full scale situations. The bench scale model should

resemble the full scale treatment plant and aerator as close as pos-

sible. All parameters such as basin geometry, mixing intensity and

turbulence, surfactant concentration, depth of tank and type of aera-

tion device should be identical to the full scale situation. Air flow

rates and KLa values are not the same for bench and full scale appli-

cations [91].

Factors Affecting a. According to Eckenfelder and Ford [92],

the a factor is a function of the constituents in the wastewater and

the turbulence regime. At high aeration intensity levels, the a coef-

ficient could be larger than unity due to entrainment of fine air bub-

bles which increases the interfacial area (A/V). Alpha increases with

increasing surfactant concentration which was attributed to decreasing

the entrained bubble size. Activated sludge suspended solids


34

concentrations up to 2000 mg/1 enhanced the rate of oxygen transfer

[89]. Synthetic suspensions were found to reduce oxygen transfer

rates at concentrations less than 2000 mg/1. Other factors

affecting a include: soluble BOD, COD, suspended solids, surface

tension, viscosity, temperature and basin configuration [91]. For a

particular aerator, alpha varies with aerator speed, the gas flow-rate

and surfactant concentration [91, 93]. Most investigators state

that a is a function of the type of aerator [72, 88, 89, 91, 93-96].

Factors Affecting 6• The beta factor, which relates the dis-

solved oxygen saturation in water to that in wastewater can be

affected by several factors which include: salts, organics, dissolved

gases, temperature, and barometric pressure [91]. Aerator speed, gas

flow rate, and surfactant concentration were also found not to have

any important effects on 6• Arora [94] states that 6 is affected by

the characteristics of the wastewater and the effects on 6 are more

significant than that of a. A 25% increase in 6 results in a 34 to

37% increase in the ratio of pounds of oxygen transferred under field

conditions to pounds of oxygen transferred under standard conditions

(N/N0 ).

Manipulation of Data. Once the oxygen transfer data are col-

lected in the field a careful analysis must be applied to ensure that

they are valid and that proper estimates for KLa and Cs (DO
35

saturation) can be determined. Several papers have been presented in

the literature describing the various methods available for analyzing

the data [93, 97-101]. These papers are directed towa~d analysis of

nonsteady state reaeration data obtained in clean tap water.

Three methods are generally available for estimating the oxygen

transfer coefficient (K1 a). The following methods include: 1) direct

or differential method, 2) exponential plot, and 3) log deficit.

Primary methods used are the log deficit and direct methods because

they are more simplistic and do not require nonlinear regression anal-

ysis. Gilbert and Chen [93] state no preference for any particular

method; however, they recommend that several methods should be used in

analysis. In their study, the use of various methods of analysis on

accurate experimental data resulted in oxygen transfer values in close

agreement and their data support the two-film theory for oxygen trans-

fer. Several investigators [97, 99-101] support the exponential

method and state that it yields the most precise estimates since data

truncation is not required. Most other studies recommend that when

utilizing other methods of analysis that truncation of data up to 20%

of the saturated DO concentration will not affect K1 a, however trunca-

tion of aeration data before reaching the saturation DO level can

produce substantial error [98, 99]. A nonlinear regression analysis

of the exponential method has been recommended to the A.S.C.E. Oxygen

Transfer Standard Subcommittee for incorporation into the proposed

standard for measurement of oxygen transfer rate in clean water

[ 85].
36

Oxygen Transfer in Biological Systems

For biological systems such as the activated sludge process to

operate efficiently, nutrients such as carbon, nitrogen, phosphorus,

trace elements, and oxygen must be supplied in the appropriate quanti-


'
ties. In many full scale plants, oxygen can be the rate limiting

nutrient. Numerous articles in the literature have reported the fail-

ure of aeration equipment under performance testing at process condi-

tions to transfer the amount of oxygen which had been specified at

standard conditions in clean tap water nonsteady state testing [72,

74, 81]. Stukenberg al. [81] have reported that only seven of

sixteen aerators passed the performance testing on the first

attempt. The aeration equipment acted as the limiting factor and

slowed down the treatment process by a factor of two to four according

to Schultze and Kooistra [i02] in their study of a full scale acti-

vated sludge treatment plant. To assure efficient treatment, oxygen

cannot be the rate limiting nutrient.

Oxygen must be in solution before it can be utilized by the sus-

pended microbial cells in the activated sludge process [42]. There-

fore oxygen must be transferred from the gas phase across a gas/liquid

interface into the bulk liquid before the microorganism can use the

oxygen. There are essentially five transport steps involved in

gas/liquid transport in biological systems: 1) diffusional transfer

(absorption) from the gas phase to the liquid medium, 2) diffusional

transfer through the bulk liquid to the cell, 3) diffusional transfer


37

across the gas film on the outer diameter of the bubble, 4) diffu-

sional transfer across the liquid film surrounding the bubble, and 5)

biochemical reaction at or inside the microorganism [42, 86, 103).

Modifications of Equation [14) to account for microbial respiration

will result in the following equation which can be used to determine

KLa in a respiring system:

[18)

R = oxygen uptake rate of the mixed liquor suspended solids,

mass/volume-time.

According to microbiologists [104-107), oxygen serves two pur-

poses in microbial systems. The primary role of oxygen is to serve as

a terminal electron acceptor for the electron transport system of

aerobic microorganisms. Oxygen is also required in small amounts as a

reactant in biochemical enzymatic reactions such as the oxidation of

hydrocarbons, synthesis of sterols and synthesis of fatty acids. It

is possible for oxygen to undergo a series of reductive reactions

inside a respiring cell and a buildup of various toxic forms of oxygen

can result. Intermediate forms of oxygen that are toxic include:

superoxide (o;), hydrogen.peroxide (H2 o2 ) and the hydroxyl radi-


1
cal (OH•) [105-107). Singlet oxygen ( o2 ) can also be produced either

chemically or biochemically. The most reactive and toxic forms are

superoxide, singlet oxygen and the hydroxyl free radical. Amino acid

oxidases form toxic products such as hydrogen peroxide [106). Aerobic

microorganisms can form enzymes such as superoxide dimutase which


38

catalyzes the reduction of superoxide to hydrogen peroxide [108).

Another enzyme, catalase decomposes hydrogen peroxide to water [106,

107]. A current article in the literature [109] states that waters

containing high organic concentrations also contain high concentra-

tions of these toxic forms of oxygen. A similar situation probably

exists in activated sludge systems. Increases in the oxygen transfer

coefficient in biological systems may in part be due to the oxygen

required to form these toxic oxygen compounds [105).

Numerous reports in the literature state that the rate of oxygen

transfer in a biological system is determined by the rate and

stoichiometry of the biological reaction [80, 96, 110]. Tsao [86]

states that oxygen transfer coefficients are larger when there is a

simultaneous biological oxidation than when there is not. Other

mechanisms have been reported for increasing oxygen transfer rates in

biological systems. Several investigators attributed increased trans-

fer rates due to direct oxygen transfer [42, 86, 103, 111, 112]. They

suggest that increased transfers result from adsorption of the micro-

bial cells onto the bubbles in the bulk liquid which results in rapid

transfer due to the overlapping of the liquid films.

The oxygen transfer rate of a particular aerator is fixed accord-

ing to standard transfer theory. Recently, two researchers [113, 114]

have alluded to KLa-being dependent on the oxygen uptake rate of the

mixed liquor suspended solids. Albertson and DiGregorio [113) have

presented data from full scale plants showing KLa increasing as the

oxygen uptake rate of the mixed liquor suspended solids increases. In


39

performance testing of low speed surface aerators treating a chemical

wastewater, Eckenfelder (114) reported the oxygenation efficiency (lb

o2 /hp-hr) increased linearly with the oxygen uptake rate. An analysis

of the data collected by Kayser [95) also shows the uptake rate

affected oxygen transfer in the same fashion. This observation could

have significant effects on conventional aerator design and testing

procedures.

SUMMARY

This review indicates there are several factors that are impor-

tant for successful operation of the completely mixed activated sludge

process. First of all, the mechanisms involved in dual substrate

limitations in the activated sludge process needs further investiga-

tion. Little research has been conducted on heterogeneous cultures

operating under multiple substrate limitations. Secondly, additional

study is required to investigate the nitrification process under dual

substrate limiting conditions to determine whether or not nitrite

accumulations may occur which would inhibit the process and lead to a

deterioration in effluent quality. Finally, several problems exist in

the aerator testing procedures currently employed. There is a need

for additional research to investigate the relationship between the

oxygen uptake rate of the mixed liquor suspended solids and the oxygen

transfer rate of the aerator in light of the results of Bennett and

Kempe [112), Albertson and DiGregorio (113) and Eckenfelder (114).


III. METHODSAND MATERIALS

The experimental materials and methods followed in this investi-

gation will now be presented. Major subheadings in this chapter

are: Experimental Approach, in which the primary objectives of the

research are stated; Laboratory Studies, in which the methods and

materials used in operating the bench scale activated sludge units are

presented; and Oxygen Transfer Studies, in which the apparatus and

methods employed in the experimental oxygen transfer study are dis-

cussed.

EXPERIMENTAL
APPROACH

Research conducted in this study can be divided into two major

areas. The first involved the determination of the biokinetic coeffi-

cients and biochemical stoichiometrical equations for the completely

mixed activated sludge process approaching operation under a dual

substrate limitation i.e. organic carbon being growth limiting at low

mean cell residence times and dissolved oxygen being growth limiting

at high mean cell residence times. Evaluation of oxygen transfer into

the process was the second area of main concern.

LABORATORY
STUDIES

.To accomplish the above objectives, two bench scale continuous

flow activated sludge reactors with cell recycle were operated over a

wide range of mean cell residence times from approximately 3 to 20

40
41

days. This range was selected since it represents high-rate to low-

rate or extended aeration treatment which are practical values uti-

lized by most full scale activated sludge treatment facilities.

The mean cell residence time (e) served as the primary control
C

parameter during the experimental study and was calculated from Equa-

tion [19]. A relatively constant value fore was obtained by wasting


C

a volume equal to the reciprocal of e times the total reactor volume


C

from the total reactor. The mixed liquor was wasted at the end of

each 24 hour operating period. This was necessary so that the total

reactor solids would stabilize over the next operating period and to

ensure that steady state conditions would exist during the next sam-

pling period.

When all of the parameters of the system remain constant over a

given time period, steady state conditions exist. In reality, steady

state operation of a biological reactor, regardless of size, is seldom

achieved. There are several factors that affect the steady state

operation of a lab scale reactor. The major factors include: fluc-

tuations in the volumetric flow rate through the reactor, experimental

errors in sampling and analytical testing, slight changes in the air

flow rate into the aeration basin, differences in ambient tempe~ature,

and population shifts of the microbial species in the mixed liquor.

The continuous flow reactors were operated over a period of

approximately three times the mean cell residence time so that steady

state conditions would exist. Steady state operation was assumed to

exist when the suspended solids concentration in the mixed liquor and
42

the effluent remained constant and the effluent pH values remained

relatively constant. Once steady state conditions were attained, the

parameters in Table I were monitored daily for a period of 7 to 10

days. The data obtained on seven consecutive days of the above time

period were then averaged together at that specified mean cell resi-

dence time.

Laboratory Apparatus

The reactors were constructed of 3/8" plexiglass and had total

operating volumes of approximately 8.5 liters. Each reactor was

divided into an aeration chamber and a clarifier by an adjustable

baffle. A slanted baffle inclined at 25° from the vertical was used

in this design to provide better mixing of the mixed liquor suspended

solids at the low air flow rates (1 liter per minute) utilized in the

study. Approximate volumes for the aeration chamber and clarifier

were 6.3 liters and 2.2 liters respectively. The treatment units were

housed in a constant temperature room which maintained a temperature

between 20° ± 1.0°c. Synthetic wastewater was pumped from 20 liter

calibrated Nalgene plastic carboys through 1/4 inch Tygon tubing by a

Calgon Model P-8 chemical feed diaphragm pump (Calgon Corporation,

Pittsburgh, Pennsylvania). The wastewater volumetric flow rate

through the reactors averaged 16.4 liters per day which resulted in

detention times for the aeration basin and clarifier of approximately

9.2 hours and 3.2 hours respectively. Each day the influent carboys

and feed lines were disinfected. A strong bleach solution was added
43

TABLE I

Parameters Monitored Daily

I. Influent Feed
A. Chemical Oxygen Demand
B. Alkalinity
C. Total Kjeldahl Nitrogen
D. NH3 -N Concentration
E. pH

II. Biological Reactors


A. Total System Microorganism Concentration
B. Aeration Basin Microorganism Concentration
C. Dissolved Oxygen Concentration
D. pH
E. Temperature
F. Steady State K1 a
G. Oxygen Uptake Rate

III. Filtered Effluent


A. Chemical Oxygen Demand
B. Alkalinity
C. Total Kjeldahl Nitrogen
D. NH3 -N Concentration
E. pH_
F. NO2 -N Concentration

G. NO3 -N Concentration

IV. Unfiltered Effluent


A. Suspended Solids Concentration
B. pH
C. Nonsteady State K1 a
D. Dissolved Oxygen Saturation Concentration
44

to each of the 20 liter influent carboys and then filled with hot tap

water. After a 30 minute disinfecting period, the carboys were rinsed

twice with hot tap water and once with cold tap water prior to use.

During this time period, a strong bleach solution was pumped through

the feed lines to inhibit biological growth. Next, the feed lines

were flushed with hot tap water for approximately 30 minutes to remove

any trace of chlorine. During the course of the study, it was neces-

sary to replace the influent lines with new 1/4 inch Tygon tubing when

the wastewater flow rate decreased due to buildup of precipitate and

biological growth in the lines.

Air for each of the reactors was introduced through two aeration

diffuser stones suspended in the back of the aeration chamber.

Throughout the study, air was supplied to each reactor at 1 liter per

minute. This provided oxygen for microbial utilization and adequate

mixing in the aeration basin to keep the mixed liquor suspended solids

in suspension. The air flow rate was controlled and measured using

Bendix rotameters (Bendix Environmental Science Division, Baltimore,

MD). Cotton filled_Nalgene tubes preceded the rotameters to prevent

contaminants from damaging the flowmeters and from entering the aera-

tion basin. Each reactor was provided with its own air source, a

Second Nature Whisper 800 aquarium air pump (Willinger Bros., Inc.,

Fort Lee, New Jersey). These aquarium pumps were utilized rather than

the in-house air since the power plants supply on campus was not

always constant nor at the same pressure.


45

Effluent from each reactor was collected in a calibrated 20 liter

Nalgene carboy. A schematic diagram of the system utilized in this

study is provided in Figure 2.

Synthetic Feed

Tables II and III list the theoretical chemical composition of

the synthetic wastewaters used in the treatability studies. Stock

solutions of the various chemicals were made and used in the daily

preparation of the synthetic feed. The theoretical chemical oxygen

demand (COD) of the feed solution was 300 milligrams per liter (mg/1)

which was provided by bacto-peptone nutrient broth. The bacto-peptone

nutrient broth also supplied a portion of the total Kjeldahl nitrogen

(TKN) requirements of the system. Nitrogen in the form of ammonium

sulfate, (NH4 ) 2 so4 , was added to the feed so that a specified COD:TKN

ratio could be achieved. Other nutrients required for growth were

supplied according to Tables II and III.

The only difference in the feed solutions to the reactors was the

TKNand alkalinity concentrations. Total Kjeldahl nitrogen concentra-

tions of approximately 50 milligrams per liter (mg/1) and 500 mg/1

were maintained in the influent to each respective reactor. This

resulted in COD:TKNratios of approximately 6.5:1 and 0.5:1 for each

of the wastewater influents. In subsequent discussions the reactor

receiving the wastewater at a COD:TKNratio of 6.5:1 will be desig-

nated as Reactor-1 (R-1). The other system where the influent COD:TKN

ratio is 0.5:1 will be called Reactor-2 (R-2). The high nitrogen


46

FEED PUMP

CALIBRATED
FEED TANK

ADJUSTABLE BAFFLE

COTTON AIR FILTER

AQUARIUM AIR
PUMP

2 DIFFUSER STONES

AERATION BASIN

._ ___ SETTLING BASIN

CALIBRATED EFFLUENT
COLLECTION TANK

Figure 2. Bench Scale Activated Sludge Unit with Internal Cell


Recycle.
47

Table II

Composition of Influent to Reactor 1

Final
Stock Quantity Concentration in
Concentration Used 18 liter
Compound (g)/2 liter (ml)/18 liter (mg/1)

Bacto-peptone
(Nutrient broth) 106.00 100.00 294.

Mgso4 • 7H2o 90.00 20.00 50.0

Mnso4 • H20 9.00 20.00 5.00

FeS0 4 • 7H20 4.00 20.00 2.22

CaC12 6.75 20.00 3.75

KCl 12~60 20.00 7.00

(NH4)2S04 398.40 4.40 48.7

K2HP04 121.34 10.00 33.7

NaHC03 *10.00 555.

*Grams of NaHco3 added directly to 18 liters of synthetic wastewater


48

Table III

Composition of Influent to Reactor 2

Final
Stock Quantity Concentration in
Concentration Used 18 liter
Compound (g)/2 liter (ml)/18 liter (mg/1)

Bacto-peptone
(Nutrient broth) 106.00 100.00 294.

MgS04 • 7H2o 90.00 20.00 50.0

MnS04 • H2o 9.00 20.00 5.00

FeS04 • 7H20 4.00 20.00 2.22

CaC12 6.75 20.00 3.75

KCl 12.60 20.00 7.00

(NH4)2S04 398.40 200.00 2213.

K2HP04 121.34 10.00 33.7

NaHC03 *108.00 6000.

*Grams of NaHC03 added directly to 18 liters of synthetic wastewater


49

content in the influent to R-2 was necessary so that a high nitro-

genous oxygen demand would exist when nitrification occurred. This

would result in a total oxygen demand that would be approximately five

times greater than that of R-1. By supplying the same amount of air

to each system, it was possible to operate R-2 under a dissolved oxy-

gen limitation at long mean cell residence times.

When nitrification occurs, the pH of the mixed liquor in the

aeration basin will decrease if sufficient alkalinity is not pre-

sent. To control the pH of the two reactors, sodium bicarbonate

(NaHco3 ) was weighed out daily and added directly to the influent

carboys when preparing the feed solution. Theoretically, 7.14 mg/1 of

alkalinity as calcium carbonate (Caco 3 ) is required per mg/1 of nitro-

gen oxidized according to stoichiometric equations. Therefore, the

total alkalinity concentrations in the influent to R-1 and R-2 was

approximately 320 mg/1 and 3600 mg/i as Caco 3 respectively. Utiliza-

tion of the NaHco3 to buffer the systems resulted in an overall influ-

ent pH of 7.8 and 7.9 to R-1 and R-2 respectively.

Every three days the bacto-peptone nutrient broth was weighed out

and dissolved in enough distilled, demineralized water to obtain one

liter of solution. The synthetic feed solution had to be prepared

daily. To ensure that the sodium bicarbonate added to the feed was

thoroughly dissolved into solution, it was first placed into the 20

liter influent carboys and 10 to 12 liters of tap water (20°C) was

added. The carboys were capped and then shaken vigorously to help

dissolve the NaHC03 • The stock solutions of magnesium and manganous


50

sulfate, ferrous sulfate, calcium chloride, potassium chloride, ammon-

ium sulfate, sodium phosphate (diabasic), and nutrient broth were

added next in that order and thoroughly mixed. More tap water was

added to dilute the feed solution so that the final total volume was

approximately 18 liters.

Acclimation and Start-up

Activated sludge for seeding the reactors was obtained from a

bench scale reactor being operated in the environmental engineering

laboratory at Virginia Polytechnic Institute and State University in

Blacksburg, Virginia. Additional activated sludge was obtained from

the effluent from the nitrification process utilized at the Roanoke

Municipal Wastewater Treatment Plant in Roanoke, Virginia to increase

the nitrifier population in the reactors.

Acclimation of the sludge was accomplished through batch type

feeding in two, 8 liter cylindrical batch reactors. The quantities of

the stock solutions and nutrient broth listed in Tables II and III

were added directly to the 8 liter batch reactors rather than diluting

them to 18 liters with tap water. During this time period, the

heterogeneous microbial culture of activated sludge became acclimated

to the substrate and the concentration of the microorganisms in the

sludge also increased.

Once the microbial solids concentration had built up to approxi-

mately 2,800 mg/1, the cultures were transferred to the 8.5 liter

continuous flow reactors. Continuous flow operation was then


51

initiated and the desired mean cell residence time was accomplished by

wasting a portion of the mixed liquor daily.

Daily Protocol

A rigorous sampling and analysis schedule was maintained once

steady state conditions were achieved. At the end of each 24 hour

period, new feed had to be made up and sludge wasting and sampling had

to be conducted prior to preparation of the synthetic feed. The

reactors were operated in such a fashion that feed was made up at 12

noon each day during the testing period. Preparation of the feed,

disinfection of influent lines and carboys, and sludge wasting took

approximately 1 hour to perform. Prior to feed preparation, it was

necessary to conduct several other analyses. The following is a brief

description of the schedule maintained during the steady state testing

period.

Two hours prior to the feeding time, the following events would

take place. First, the temperature of the mixed liquor was recorded

for R-1 and R-2. This was followed by a calibration of the dissolved

oxygen probe against moist air. Once the probe was calibrated, the

dissolved oxygen concentration was determined in the aeration

basins. Then 10 to 20 milliliters (ml) of mixed liquor from the aera-

tion basin was collected for total suspended solids determination.

The last step involved determination of the oxygen uptake rate of the

mixed liquor suspended solids in the aeration basin.


52

At each feeding time, the next sequence of events occurred.

First, the influent pumps were turned off. Effluent lines from each

reactor were clamped and the adjustable baffle was removed to estab-

lish complete mixing of the mixed liquor suspended solids. The influ-

ent lines and carboys were then disinfected as previously outlined. A

specified amount of mixed liquor was siphoned from the reactor to

maintain the desired mean cell residence time. A portion of the

wasted sludge would be utilized for determination of the total system,

total suspended solids concentration and pH. The baffle was then

reinserted and the sludge allowed to settle to the bottom of the clar-

ifier. After this was accomplished, a sample of approximately 800 ml

was obtained from each calibrated, effluent carboy. The remaining

effluent was transferred to two 20-liter Nalgene carboys for storage

until the oxygen transfer studies could be conducted in the after-

noon. The calibrated effluent carboys were then rinsed once with hot

tap water and were then ready to be used for effluent collection dur-

ing the next 24 hour operating period. Two hundred milliliters of

each effluent sample were utilized for suspended solids determina-

tion. The remaining 600 ml of sample was ut{lized for pH determina-

tion and then filtered through a 0.45 µ (micrometer) filter. Approxi-

mately 250 ml of each filtered effluent was acidified with concen-

trated sulfuric acid (H2 so4 ) to a pH of about 2 and then stored at

4°c. These samples were utilized at a later date for chemical oxygen

demand (COD), total Kjeldahl nitrogen (TKN) and ammonia nitrogen

(NH3-N) determinations. One hundred milliliters of each filtered


53

effluent was collected and frozen to be utilized for nitrate

nitrogen (No;-N) determinations. The remaining filtered effluent from

each reactor was utilized immediately after preparing the synthetic

feed for alkalinity and nitrite nitrogen (No;-N) determinations. For

the alkalinity testing, 100 ml of filtered effluent was utilized from

R-1 and 10 ml of filtered effluent diluted to 100 ml with distilled

water was utilized from R-2.

Each day the synthetic wastewater had to be prepared according to

Tables II and III. After the wastewater was prepared, a pH probe was

lowered into each influent carboy for pH determination. At this time,

50 ml of sample from the influent to R-1 and 10 ml of sample from the

influent to R-2 was collected and diluted to 100 ml with distilled,

demineralized water for alkalinity determination. Influent samples of

250 ml were collected and acidified with concentrated H2 so4 to a pH of

approximately 2. These samples were utilized at a later date for COD,

TKN, and NH3-N determinations.

Analytical Procedures

The analytical techniques used in this study are as follows:

Chemical oxygen demand. The chemical oxygen demand of the syn-

thetic feed solutions and soluble or filtered effluent samples was

determined by the dichromate reflux method as outlined in Standard

Methods for the Examination of Water and Wastewater [80]. The CODof

soluble filtered effluent samples containing nitrite nitrogen (No;-N)


concentrations greater than 1-2 mg/1 was determined as_ above except
54

that 10 milligrams (mg) of sulfamic acid per mg of nitrite nitrogen

was added to the dichromate solution as specified in Standard Methods

[80].

Temperature. Temperature was recorded to the nearest o.s 0 c using

a mercury filled thermometer as described in Standard Methods [80].

pH. A Fisher Accumet Model 120 pH meter (Fisher Scientific

Company) was utilized for pH determinations. The hydrogen ion concen-

tration or pH was determined by the Glass Electrode Method described

in Standard Methods [80].

Suspended solids. Suspended solids concentration was determined

as total nonfiltrable residue dried at 103-105°c as described in

Standard Methods [80] using a 0.45 µ Whatman Glass Microfibre filter

(Whatman Laboratory Products, Inc., Clinton, NJ).

Total Kjeldahl nitrogen. As described in Standard Methods [80],

the total Kjeldahl nitrogen was determined by a digestion/distillation

procedure with an acidimetric titration finish. The macro-

digestion/distillation procedure was utilized for all of the analyses

except for the samples collected at the first mean cell residence time

in which the micro-digestion/distillation procedure was used.


55

Ammonia nitrogen. Ammonia nitrogen (NH3-N) concentration of the

feed solutions and soluble effluent samples was determined by the

micro-distillation procedure as described in Standard Methods [80].

Nitrite nitrogen. Soluble effluent samples were analyzed for

nitrite nitrogen (No;-N) concentration by the diazotization method

utilizing sulfanilic acid and N-(1-napththyl)-ethylenediamine

dihydrochloride as described in Standard Methods [80].

Nitrate nitrogen. The nitrate nitrogen (No;-N) concentration of

the soluble effluent samples was determined by the brucine method as

described in the United States Environmental Protection Agency's

Manual of Methods for Chemical Analysis of Water and Wastes [115].

Dissolved oxygen. Dissolved oxygen (DO) concentration was mea-

sured with a YSI Model 54-ARC Oxygen Meter (Yellow Springs Instrument

Company, Inc., Yellow Springs, Ohio). The instrument was calibrated

daily by reading against air according to the manufacturer's instruc-

tions as specified in Standard Methods [80].

Oxygen uptake rate. The oxygen uptake rate or oxygen consumption

rate of the mixed liquor suspended solids was determined by the proce-

dure described on pages 125-126 of Standard Methods [80].


56

Alkalinity. Samples of the synthetic feed solutions and soluble

effluent were analyzed for total alkalinity in mg/1 as Caco 3 • A

potentiometric titration to a pH of 4.5 as described in Standard

Methods [80] was employed for the alkalinity determinations.

OXYGEN
TRANSFERSTUDIES

Separate studies were conducted to investigate the transfer of

oxygen into the completely mixed activated sludge process. The over-

all oxygen transfer coefficient (K1 a) was determined from steady state

tests conducted on activated sludge and from nonsteady state tests

using wastewater effluent.

To determine the K1 a values under steady and nonsteady state

testing conditions, it was necessary to operate the two bench scale

activated sludge units described previously to generate activated

sludge and wastewater effluent at selected mean cell residence

times. The laboratory apparatus required for the operation of these

units was previously discussed.

Laboratory Apparatus

Determination of the K1 a coefficient under nonsteady state condi-

tions required the construction of a separate plexiglass vessel in

which to conduct the studies. An 8 liter cylindrical unit was con-

structed of 1/4 inch plexiglass. At the bottom of the vessel, an air

inlet was made so that oxygen could be transferred into the unit and

released through a porous diffuser stone. The container was fitted


57

with a top which housed an electric motor (WAMCO


Corporation) support-

ing a 14 1/2 inch aluminum shaft to which a 3 inch diameter impeller

was fitted for mixing of the contents of the reactor. A Powerstat-

variable Autotransformer (Superior Electric Company, Bristol,

Connecticut) was utilized in controlling the degree of mixing required

in the reactor.

In-house air was supplied to the unit during test periods at a

rate of 1 liter per minute. All tests were conducted in a constant

temperature room at 20°c ± l.0°c. The air flow rate was measured and

controlled by a Bendix rotameter (Bendix Environmental Science

Division, Baltimore, MD). A cotton filled Nalgene plastic tube pre-

ceded the rotameter to avoid oil and other contaminants from damaging

the flow meter and from entering the reaction vessel.

Nonsteady State Testing Procedure

Nonsteady state KLa values were obtained in separate batch tests

conducted on the effluent from each of the reactors during the labor-

atory studies. The alpha (a) and beta (e) coefficients were also

determined for the synthetic wastewater at each of the mean cell resi-

dence times studied.

The procedure for determining the nonsteady state KLa values

involved deoxygenation of the wastewater followed by a period of

reaeration as described in the Environmental Engineering Unit

Operations and Unit Processes Laboratory Manual (116]. Six liters of

effluent were added to the KLa apparatus for testing purposes. The
58

temperature and dissolved oxygen concentration were determined accord-

ing to Standard Methods [80]. The third step involved the addition of

sodium sulfite (Na 2 so3 ) for deoxygenating the water and cobaltous

chloride (CoC12 °6H2 0) was added to the reaction vessel to catalyze

this reaction. Next, the motor was turned on and air was injected

into the vessel at 1 liter per minute. The dissolved oxygen concen-

tration was recorded as a function of time until the dissolved oxygen

saturation value had been reached. A Fisher LCD Digital stopwatch was

used to time the experiment. Oxygen saturation was assumed to exist

when the DO reading was the same for three consecutive readings. At

this time the motor and air were turned off and testing discon-

tinued. The unit was then drained and rinsed thoroughly with hot

water and again with cold tap water before the next run.

The same procedure utilizing Blacksburg, Virginia tap water was

followed to determine the overall oxygen transfer coefficient in clean

water. This was necessary in order to determine the alpha and beta

coefficients for the synthetic wastewater.

Steady State Testing Procedure

Determination of the overall oxygen transfer coefficient (KLa) at

steady state conditions was accomplished by operating two bench scale

activated sludge units as outlined in the section on laboratory stu-

dies. It was possible to calculate the KLa value directly by knowing

the dissolved oxygen (DO) concentration in the aeration basin, the

dissolved oxygen saturation concentration for the wastewater at the


59

specified temperature, and the oxygen uptake rate of the mixed liquor

suspended solids. All of these parameters were collected during the

laboratory studies on the synthetic wastewater.

The procedures used for determining the dissolved oxygen concen-

tration in the aeration basin and the oxygen uptake rate of the mixed

liquor were outlined in the anaytical procedures sectfon under labor-

atory studies. Oxygen saturation values for the synthetic wastewater

were obtained during the nonsteady state tests.


IV. MATHEMATICAL
MODELING

Various mathematical descriptions have been proposed to model the

completely mixed activated sludge process [2-10]. Perhaps the most

widely known is the unified theory developed by Lawrence and McCarty

[2]. This model has been enhanced by Sherrard and coworkers [9, 117]

by the addition of biochemical stoichiometric equations which can be

used to predict various parameters in the activated sludge process.

Successful utilization of the model requires a thorough labor-

atory investigation to determine the biokinetic coefficients that must

be used in the equations. Normally, two sets of biokinetic constants

are required, one for the heterotrophic reaction and a second set for

the nitrification reaction. The nitrification reaction (Equations [1]

and [2]) was previously shown to be a two-step sequential reaction.

Therefore, it is possible to determine the biokinetic constants for

each individual reaction, the ammonium oxidation reaction involving

Nitrosomonas and the nitrite oxidation reaction involving

Nitrobacter.

After obtaining these three sets of kinetic data, predictions of

effluent substrate concentration, percent removal efficiency, nutrient

requirements, waste sludge production, aeration basin mixed liquor

suspended solids concentration, observed yield coefficient, and alka-

linity production or destruction can be made. Each of these par-

ameters are specific for a given mean cell residence time, (e ).


C

60
61

BIOKINETICEQUATIONS

The mean cell residence time (e) represents the average time a
C

unit of biomass remains in the reactor. This is the primary parameter

used to control the degree of treatment in the activated sludge pro-

cess. Mean cell residence time is calculated from the following rela-

tionship:

[19]

where ec = mean cell residence time, time,

V = aeration basin volume, volume,

X = aeration basin microorganism concentration,

mass/volume,

XE= effluent liquid microorganism concentration,

mass/volume,

~=wastage microorganism liquid flow rate, volume/time,

and

Q = effluent liquid flow rate, volume/time.


E
Effluent substrate concentration can be determined from the fol-

lowing equation:

[20]

where: s1 = effluent substrate concentration, mass/volume,

Ks= substrate concentration at one-half the maximum rate of

substrate utilization per unit weight of micro-

organisms, mass/volume,
62

k = maximum rate of substrate utilization per unit weight

of microorganisms, time- 1 ,

kd = microorganism maintenance or decay coefficient, time- 1 ,

and

Ymax = maximum microorganism yield coefficient,

mass/mass.

For any particular substrate constituent, the percent removal

efficiency can be determined in the following manner:

E [21]
s

where: Es= efficiency of substrate removal, percentage,

S0 = influent substrate concentration, mass/volume, and

s1 = effluent substrate concentration, mass/volume.

An observed yield coefficient which represents the actual unit of

biomass produced per unit of substrate utilized can be calculated with

the following equation:

y
max
[22]

where: Yobs= observed microorganism yield coefficient,

mass/mass.

Mixed liquor suspended solids concentration in the aeration basin

can be determined as follows:


63

Ymax (So - Sl) 8c


X = [23]
(1 + kd .ec) 8

where: e = hydraulic retention time, time.

At steady state conditions, the amount of biomass generated is

equal to the amount that must be wasted from the system. Waste sludge

production can be determined by solving the following expression:

[24]

where: Px = waste sludge production per unit time, mass/time, and

Q = influent liquid flow rate, volume/time.

STOICHIOMETRICEQUATIONS

The use of the standard biokinetic equations [2] along with

stoichiometric relationships [6, 9, 10] can be used to develop

balanced biochemical stoichiometric relationships which can be used to

model the completely mixed activated sludge process. Quantitative

stoichiometric equations can be derived from qualitative

stoichiometric equations for any given mean cell residence time.

Biokinetic and stoichiometric relationships can be used for predicting

effluent substrate concentration, mixed liquor suspended solids con-

centration in the aeration basin, nutrient requirements, waste sludge

production and oxygen requirements.


64

Qualitative Reactions

Qualitative equations can be written to represent the heter-

otrophic and nitrification reactions that occur in the activated

sludge process. The following word equation represents the heter-

otrophic reaction.

Substrate+ Bacteria+ Nutrients+ Oxygen-> New Bacteria+ Water

+ Carbon Dioxide+ Residual Organics and Inorganics [25]

A qualitative expression for the autotrophic reaction can be written

as follows.

Ammonia+ Bacteria+ Oxygen-> New Bacteria+ Nitrate

+ Residual Nitrogen [26]

Quantitative Reactions

As stated previously, quantitative stoichiometric equations can

be written to represent the process and the nitrification reaction can

be replaced with a two-step sequential stoichiometric equation. The

following are the biochemical stoichiometric equations for the com-

pletely mixed activated sludge process with cellular recycle utilizing

glucose (C 6H12 o6 ) for the organic carbon source and ammoniwn sulfate

as the nitrogen source. The following equation represents the heter-

otrophic or COD removal reaction. Elemental chemical composition of

the microbial biomass is represented by the formula c5H7 o2N and has
65

been shown to remain relatively constant over a wide range

of eC values .[118].

C6H12o6 + z 3 NH!
(27]

A quantitative equation for ammonium oxidation can be written as fol-

lows.

x 4 CO2 + z 3 NH1
+ x5 0 2 -) z 7 C5H70 2N + z 8 NH:

+ z 9 N0-2 + z 10 H+ + z 11 H2 o (28]

The nitrite oxidation reaction can be expressed quantitatively as

follows.

x 6 CO2 + z 9 N02 + x 7 H20 + x 8 o2 + x 9 H+ -) z 12 C5H70 2N

+_z13 N02 + z14 N03 [29]

Summation of the three previous equations result in the overall reac-

tion for the process .•


66

Carbon Limitations

A knowledge of the influent characteristics and the predicted

effluent characteristics determined from biokinetic equations are used

in developing the biochemical stoichiometric equations. Solutions to

the coefficients for each constituent can be accomplished in the fol-

lowing manner when organic carbon is the only growth limiting nutri-

ent.

Coefficients x 1 and x 2 can be determined by knowing the influent

characteristics and composition such as CODand NH3-N. The effluent

glucose coefficient z 1 is set by the effluent glucose concentration

predicted by the biokinetic equations. z 1 is determined from the

following expression:

[31]

The coefficient for effluent ammonium nitrogen (NH:-N) is determined


by a mass balance on nitrogen and z 4 is calculated by a mass balance

on carbon. A charge balance must exist for the equation to balance,

therefore, z 6 can be determined knowing that x2 = z 3 - z6 • Performing

mass balances on hydrogen and oxygen will result in the values for z5

and x 3 •

If the mean cell residence time is sufficiently long enough that

nitrification is occurring, stoichiometric equations can be developed

for ammonium oxidation and nitrite oxidation (Equations [l] and

[2]).
67

The coefficients for ammonium oxidation are solved as follows.

Influent ammonium nitrogen is the same as that for NH+


4 -N in the efflu-
ent from the heterotrophic reaction. Effluent NH+-N concentration
4
predicted by the biokinetic equation fixes the value for z8 • Next, z7

can be determined from the following equation:

A mass balance on nitrogen will result in the solution for z9 • Since

a charge balance must exist, z 10 = z 3 - z8 + z9 • Conducting mass

balances on carbon, hydrogen, and oxygen will yield solutions to x 4 ,

z 11 , and x 5 respectively.

The numerical coefficient for the N02 in the influent fo~ the

Nitrobacter reaction is the same as that for N02 in the effluent from

the Nitrosomonas reaction. z 13 is set by the effluent NO; concentra-

tion determined by the biokinetic equations. Determination of z 12 is

accomplished by solving the following relationship:

[33]

Performing mass balances on nitrogen and carbon will result in values

for z 14 and x 6 • The x 9 coefficient is determined from a charge

balance knowing that x 9 = z 9 - z 13 - z 14 • All remaining coefficients

are determined by balancing the equati~n in the usual manner. A sum-

mation of Equations [27], [28], and [29] results in the overall

stoichiometric equation for the total treatment reaction (Equation

[30]).
68

Oxygen Limitations

The procedure developed above is only appropriate when carbon is

the growth limiting nutrient. When aeration capacity is exceeded, an

alternate procedure must be utilized since oxygen becomes the limiting

substrate. Dissolved oxygen concentration effects on the hetero-

trophic-autotrophic reactions have been previously discussed. Accord-

ing to Focht and Verstraete [34] heterotrophs out-compete the auto-

trophs for oxygen. It has also been established that ammonium oxi-

dizers out-compete the nitrite oxidizers to the point of completely

inhibiting them [35]. Stoichiometric equations developed under oxygen

limitations assumes that the oxygen requirements for the three reac-

tions are satisfied in the following order: heterotrophs, ammonium

oxidizers, and nitrite oxidizers. When a plant's aeration capacity is

exceeded due to an overload either organically and/or hydraulically,

the amount of air transferred may be sufficient to satisfy all the

carbonaceous oxygen demand and only a portion of the oxygen required

for complete ammonium oxidation. In such instances, an accumulation

of nitrites could occur leading to inhibition of nitrification and a

deterioration in effluent quality.

Since oxygen utilization by the heterotrophs takes precedence

over autotrophic utilization and all the carbonaceous oxygen demand is

assumed to be satisifed, the stoichiometric equations for the hetero-

trophic reaction are developed as outlined in the previous section.

For the ammonium oxidation however, only a portion of the oxygen

required is available due to limited aeration capacity. The


69

Nitrosomonas reaction can be developed mathematically on a mathe-

matical proportional basis. A proportion of the actual amount of

oxygen transferred to the ammonium oxidizers to the total amount of

oxygen required for complete ammonium oxidation is computed. This

value is used to multiply the following coefficients z 7 , x 4 , z 9 , z 10 ,

and z 11 (which were obtained under carbon limiting conditions) in

order to obtain the new values for each coefficient under oxygen

limiting conditions. The coefficient for effluent NH+


4 -N from the
heterotrophic reaction is the same as the coefficient for

influent NH!-N for the ammonium oxidation reaction. Performing a mass

balance on nitrogen, will result in a solution for z 3 • Nitrite oxida-

tion is completely inhibited, therefore all the following coefficients

(x 6 , x7 , x 8 , x 9 , z 12 and z 14 ) are set equal to zero except for z 13

which is the same as z 9 • A similar procedure can be utilized if the

oxygen requirements for both the heterotrophic and ammonium oxidation

reactions are sattsfied but only a portion of the oxygen requirements

are met for the nitrite oxidation reaction.

SUMMARY
Knowing the influent wastewater characteristics and the bio-

kinetic constants for each of the reactions, it is possible to predict

the concentration of any particular constituent desired. These equa-

tions can be easily developed into a Fortran program for solving and

the results generated can be output as graphical plots to show how

each constituent varies with the mean cell residence time. A listing
70

of the Fortran and SAS programs used in the theoretical modeling study

can be found in Appendix A.


V. RESULTS

Experimental and theoretical results derived from this study are

contained in the following presentation. This chapter is divided into

the following sections: Experimental and Theoretical Results, in

which the laboratory and theoretical data are presented and Oxygen

Transfer Study Results, in which the data collected in the experi-

mental oxygen transfer study are presented.

To accomplish the objectives of this research, it was necessary

to operate two continuous flow, completely mixed activated sludge

units for approximately one year. The details of materials and

methods employed in this study have been described in Chapter III.

After experiencing numerous difficulties with proper influent sub-

strate selection and regulating feed pumping rates, data were col-

lected on each laboratory reactor for approximately nine months.

The units were operated so that Reactor-1 (R-1) would always be

growth limited with respect to carbon and Reactor-2 (R-2) would be

carbon limiting at low mean cell residence times and oxygen limiting

at high sludge ages. Each system was supplied the same amount of air

and the influent CODwas essentially the same, approximately

330 mg/1. COD:TKNratios of R-1 and R-2 averaged 6.07:1 and 0.65:1

respectively. A low COD:TKNratio for R-2 was required so that a high

nitrogenous oxygen demand would be exerted at high mean cell residence

times. Theoretically, the influent to R-2 was to have a total oxygen

demand of approximately five times that of R-1 so that oxygen would be

71
72

the limiting nutrient in R-2 at high eC values. Investigation of dual

substrate limitations in the activated sludge process required this

type of operation.

The data presented at each mean cell residence time represent

daily data values averaged over at least a seven day period where

steady state conditions existed. Most of the data are presented in

graphical form utilizing the elemental distribution diagram approach

as described by Sherrard and Benefield [119]. Raw data for each of

the five experimental runs for each reactor are presented in tabular

form in Appendix B.

A theoretical study was conducted in which a mathematical model

was developed using biokinetic and stoichiometric equations to simu-

late the completely mixed activated sludge process operating under

carbon and oxygen limitations. Development of the Fortran computer

program used to simulate the process was discussed in Chapter IV

entitled "Mathematical Model". Both Reactor 1 and 2 were modeled in

the theoretical study. During the experimental study, R-1 was oper-

ated under a carbon limitation and air was supplied in excess. In the

theoretical study, simulation of R-1 operating under a carbon limita-

tion rather than an oxygen limitation required the amount of oxygen

transferred to the process to be set at 100 millimoles per day so that

oxygen would not restrict growth of the microorganisms. To test the

model approaching a dual substrate limitation, the amount of oxygen

transferred to the process was restricted to 56.7 millimoles of oxygen

per day which resulted in carbon being the rate limiting nutrient at
73

low mean cell residence times and oxygen being limiting at sludge ages

greater than or equal to 4.3 days. At higher mean cell residence

times, the nitrite oxidation reaction was only supplied a portion of

the total oxygen requirements necessary for the complete oxidation of

nitrite to nitrate and at a 8


C
= 16.5 days this resulted in complete

inhibition of the Nitrobacter. This second situation was investigated

to determine if a nitrite and/or nitrate accumulation could be pre-

dicted by using the model.

A sensitivity analysis was conducted to determine the autotrohpic

biokinetic coefficients which would result in the experimental results

matching the results predicted by the model. Several values and var-

ious combinations of the coefficients were utilized to get the best

fit of the data. The sensitivity analysis was_only conducted on

Reactor 1 since it operated under a carbon limitation and was not

inhibited due to high concentrations of free ammonia or nitrite as

experienced in R-2 during the laboratory studies. All values for the

heterotrophic biokinetic constants were held constant during the sen-

sitivity study. The values used for Ymax and kd were derived from the

experimental data collected on R-1 and the values fork and Ks were

obtained from the literature [49]. Table IV contains the final bio-

kinetic constants used in the mathematical model for the heterotrophic

and autotrophic reactions. Other parameters used in the modeling

study were obtained from actual operating data during the laboratory

investigation. These values are tabulated in Table V.


74

Table IV

Biokinetic Coefficients Used in Mathematical Model

Treatment Objective Coefficient Value Basis

ymax 0.383 1 mg MLSS/mg COD

CODRemoval kd 0.057 1 1/day

k 6.000 2 1/day

Ks 120.000 2 mg/1 COD

+
ymax 0.150 mg MLSS/mg NH4 -N

AmmoniumOxidation kd 0.050 1/day

k 4.000 1/day

Ks 5.000 mg/1 NH+-N


4

ymax 0.050 mg MLSS/mg No;-N

Nitrite Oxidation kd 0.010 1/day

k 6.000 1/day

Ks 1.000 mg/1 NO;-N

1values obtained from experimental study


2 After Sherrard [49]
All other coefficients after Lawrence and McCarty [2]
75

Table V

Mathematical Model Parameters

Reactor Parameter Value Basis

Influent COD 334.0 mg/1

Influent NH+-N 55.0 mg/1


4
R-1 Volumetric Flowrate 16.7 1/day

Hydraulic Retention Time 0.395 day


+
COD:NH
4-N Ratio 6.07:1 mg/1 COD/mg/1 NH+-N
4
Sludge Age 3.9-21.0 day

Influent COD 322.0 mg/1

Influent NH+-N 493.1 mg/1


4
R-2 Volumetric Flowrate 16.0 1/day

Hydraulic Retention Time 0.419 day

COD:NH:-N Ratio 0.65:1 mg/1 COD/mg/1 NH+-N


r
Sludge Age 2.2-16.5 day
76

In the sensitivity analysis of Reactor 1, the following trends

were observed when the values fork and Ks were varied. Only k and Ks

were varied for each of the autotrophic reactions while Ymax and kd

were held constant. Values selected from the literature for Ymax and

kd were 0.15 and 0.05 for ammonium oxidation and 0.05 and 0.01 for

nitrite oxidation [2]. Increasing the magnitude of k for the ammonium

oxidation reaction resulted in the model predicting higher values for

both No;-N and N03 -N at each specified mean cell residence time. For

ammonium oxidation, increasing Ks resulted in the values for N02N

remaining the same and the values for'No;-N decreasing. The model

predicted lower values for N02 -N and higher values for No;-N at each

mean cell residence time when the magnitude of k was increased for the

nitrite oxidation reaction. Decreasing K6 for the Nitrobacter reac-

tion resulted in lower values for N02 -N and higher values for NO;-N at

the specified eC value.

The biokinetic coefficients determined from the sensitivity

analysis on R-1 were used in the simulation of R-2 since the waste-

water characteristics were essentially the same except the influent

nitrogen concentration to R-2 was much higher. This should have

resulted in the model's predictions matching the experimental data for

R-2, however, there were high concentrations of nitrite and free

ammonia present in the mixed liquor of R-2 which inhibited the process

during the experimental studies. Refinement of the model to account

for i~hibition would result in a better model.


77

Substrate inhibition models as proposed by Andrews and Graef

[120] and Keenan~ al. [65] which are functions proposed by Haldane

[121] could be used to account for substrate and product inhibition.

The inhibition function proposed by Haldane is pH dependent and there-

fore it is possible to determine the concentration of unionized as

well as ionized substrate present. Haldane's model [1°21] was based on

the inhibition of enzymes at high substrate concentrations and can be

expressed as follows:

[34]

where: µ=specific growth rate, day- 1 ,

µ=maximum. specific growth rate, day- 1 , and

k. = inhibition constant, ~oles/liter,


1

and all other parameters have been previously defined. Inhibition

functions similar to Equation [34] were developed to account for sub-

strate inhibition to nitrite and free ammonia. These functions were

unsuccessfully incorporated into the Fortran computer program. Since

validation of the Lawrence and McCarty [2] model was not a primary

objective of this research, no further time or money was allocated to

refine the model with regards to substrate inhibition.

A major assumption of the model is that glucose serves as the

organic carbon source but in the experimental study, bacto-peptone was

actually used. Influent ammonium nitrogen concentration used in the

model represented the total Kjeldahl nitrogen concentration used in


78

the laboratory study. These two assumptions should not significantly

affect the results of the model since the TKN of the actual synthetic

wastewater consisted primarily of NH+


4 -N (460 mg/1 as N) and bacto-
peptone contained only a small amount of organic nitrogen, approxi-

mately 33 mg/1 as N.

EXPERIMENTAL
AND THEORETICALRESULTS

A series of tables and graphs will be used to present the results

of the laboratory and modeling data obtained in this investigation.

The results of the experimental and theoretical studies are presented

for each reactor studied. Tables VI and VII provide summaries of the

steady state experimental data and computer generated data on R-1.

Experimental and theoretical data obtained on R-2 are tabulated in

Tables VIII and IX.

In the experimental studies conducted on R-1, steady state condi-

tions were maintained at sludge ages of 3.9, 4.2, 8.5, 13.8 and 21.0

days and these were used in the theoretical study. Five steady state

conditions were investigated at mean cell residence times of 2.2, 4.3,

7.9, 15.7 and 16.5 days for R-2 during the laboratory study. These

sludge ages were employed in the computer modeling investigation.

Chemical Oxygen Demand

For R-1, influent CODconcentration and CODremoval efficiency

averaged 334 mg/1 and 93.9% during the course of the experimental

study (Table VI). Average influent wastewater COD concentration and


Table VI

Summary of Steady State Data for COD:TKN= 6.07:1

Value for Given eC , days

Parameter 3.9 4.2 8.5 13.8 21.0

COD:TKNratio 6.29:1 6.35:1 5.52:1 5.89:1 6.45:1

COD
Feed (mg/1) 343 346 321 324 336
Effluent (mg/1) 30 27 24 19 18 -..J
Removal efficiency (%) 91.3 92.2 92.5 94 .1 94.6
\0

TKN concentration
Feed (mg/1) 55.0 54.5 58.2 55.0 52.1
Effluent (mg/1) 15.1 7.5 4.4 4.2 .09
Removal efficiency (%) 72.5 86.2 92.4 92.4 99.8
NH3-N concentration
Feed (mg/1) 10.9 10.5 13 .1 9.8 10.1
Effluent (mg/1) 12.6 5.2 0.5 2.3 0.3
Removal efficiency (%) -13.5 50.5 96 .2 76.5 97.0

NO~-N concentration
ffluent (mg/1) 6.7 1.1. 2.2 0.4

NO~-N concentration
ffluent (mg/1) 29.9 40.6 45.1 43.0 50.8
Table VI (cont'd)

Summary of Steady State Data for COD:TKN= 6.07:1

aC ,
\
Value for Given days

Parameter 3.9 4.2 8.5 13.8 21.0

Biological solids
Total system (mg/1) 842 1119 1455 2032 2118
Mixed liquor (mg/1) 973 1.160 1677 2331 2432
Effluent (mg/1) 15.2 26.5 4.9 6.8 2.6
00
0
pH
Feed 7.9 8.0 7.7 7.9 7.7
Mixed liquor 7.5 7.5 7.2 7.4 7.4
Effluent 7.6 7.7 7.4 7.6 7.5
Dissolved Oxygen
Mixed Liquor 3.60 2 .10 3.90 3.50 3.65
Temperature
Mixed Liquor 20.0 20.5 20.0 20.5 19.5
Alkalinity as Caco 3
Feed (mg/1) 392 386 374 367 383
Effluent (mg/1) 281 271 186 185 196
Table VII

Theoretical Data Generated for COD:TKN= 6.07:1

Value for Given eC , days

Parameter 3.9 4.2 8.5 13.8 21.0

+ ratio
COD:NH 6.07:1 6.07:1 6.07:1 6.07:1 6.07:1
4
COD
Feed (mg/1) 334 334 334 334 334
Effluent (mg/1) 19 18 10 7 6 (X)

Removal efficiency (%) 94.3 94.6 97.0 97.9 98.2 I-'

NH3-N
Feed (mg/1) 55.0 55.0 55.0 55.0 55.0
Effluent (mg/1) 5.2 4.6 1.9 1.3 1.0
Removal efficiency(%) 90.5 91.6 96.5 97.6 98.2

NO -N
tffluent (rng/1) 7.9 4.8 0.7 0.4 0.2
NO -N
~£fluent (mg/1) 28.9 32.7 41.2 43.9 46.0
Biological solids
Total mixed liquor (mg/1) 1034 1106 1936 2654 3325
Alkalinity
alkalinity
as Caco
(mg 1) 7 -309 -314 -339 -350 -358
Table VIII

Summary of Steady State Data for COD:TKN= 0.65:1

Value for Given eC , days

Parameter 2.2 4.3 7.9 15.7 16.5

COD:TKNratio 0.68:1 0.69:1 0.60:1 0.68:1 0.62:1

COD
Feed (mg/1) 339 342 291 333 304
Effluent (mg/1) 70 55 41 24 32 00
N
Removal efficiency (%) 79.4 83.9 85.9 92.8 89.5

TKN concentration
Feed (mg/1) 498.9 498.8 488.8 491.1 488.5
Effluent (mg/1) 450.5 331.4 263.2 261.2 156.4
Removal efficiency (%) 9.6 33.6 46.2 46.8 68.0
NH3-N concentration
Feed (mg/1) 462.0 452.3 462.0 458.7 467.0
Effluent (mg/1) 441.8 330.9 261.6 256.9 161.3
Removal efficiency (%) 4.4 26.8 43.4 44.0 65.5
NOt-N concentration
ffluent (mg/1) 23.5 133.9 181.0 172.6

NO~-N concentration
ffluent (mg/1) 3.8 14.4 16.9 42.3 31.0
Table VIII (cont'd)

Summary of Steady State Data for COD:TKN= 0.65:1

Value for Given aC , days

Parameter 2.2 4.3 7.9 15.7 16 .5

Biological solids
Total system (mg/1) 427 916 2097 3362 2935
Mixed liquor (mg/1) 479 1061 2535 3226 3585
Effluent (mg/1) 49.9 31.4 22.4 25.0 30.2
00
w
pH
Feed 8.0 8.0 7.8 8.0 7.8
Mixed liquor 8.5 8.2 8.0 8.2 7.9
Efflue~t 8.6 8.4 8.2 8.3 8.0
Dissolved oxygen
Mixed liquor 5.65 2.90 .85 .90 .35
Temperature
Mixed liquor 20.0 20.0 20.0 20.0 20.0
Alkalinity as Caco3
Feed (mg/1) 3557 3590 3543 3521 3531
Effluent (mg/1) 3400 2704 2195 2245 1689
Table IX

Theoretical Data Generated for COD:TKN= 0.65:1


(Nitrite Oxidation Inhibited)

Value for Given ec, days

Parameter 2.2 4.3 7.9 15.7 16.5

COD:TKNratio 0.65:1 0.65:1 0.65:1 0.65:1 0.65:1

COD
Feed (mg/1) 322 322 322 322 322
Effluent (mg/1) 34 17 10 7 6 00
.,:-.
Removal efficiency (%) 89.4 94.7 96.9 97.8 98.1
NH3-N
Feed (mg/1) 493.1 493.1 493.1 493.1 493.1
Effluent (mg/1) 26.4 4.5 2.1 1.2 1.1
Removal efficiency (%) 94.6 99.1 99.6 99.8 99.8

NO -N
~£fluent (mg/1) 446.9 359.8 411.7 474.6 479.3
NO -N ·
~£fluent (mg/1) o.o 109.3 62.4 4.3 o.o
Biological solids
Total mixed liquor (mg/1) 836 1619 2579 3916 4016
Alkalinity
alkalinity
as Caco
7
(mg 1) -3263 -3420 -3447 -3468 -3469
85

CODremoval efficiency were 322 mg/1 and 86.3% respectively for R-2

(Table VIII). Experimental and theoretical CODremoval efficiencies

are plotted as a function of mean cell residence time in Figures 3-a

and 3-b and were found to increase as thee value increased for each
C

COD:TKNratio studied. Figures 4-a and 4-b illustrate the effect of

mean cell residence time on the effluent COD concentration. Effluent

CODconcentration increased as the sludge age decreased for both the

theoretical and experimental studies.

Mixed Liquor Suspended Solids Concentration

Graphical representations of the total reactor microorganism

concentration in the aeration basin versus the mean cell residence

time are plotted in Figures 5-a and 5-b. Actual total reactor micro-

organism concentration in the aeration basin of R-1 increased from 973

mg/1 ate = 3.9 days to 2432 mg/1 ate = 21.O days as shown in
C C

Figure 5-a. Total reactor mixed liquor suspended solids concentra-

tions ranged from 479 mg/1 to 3585 mg/1 for the experimental study on

R-2 at the corresponding sludge ages of 2.2 and 16.5 days. The the-

oretical total microorganism concentration in the aeration basin pre-

dicted by the model is plotted as a function of e in Figure 5-b.


C

TKN and NH3-N Removal Efficiences

Both TKN and NH3-N removal efficiencies increased as the mean

cell residence time increased as illustrated in Figures 6, 7-a and

7-b. The actual and predicted ammonia nitrogen removal efficiencies


86

• •
100 --~.--~.---,----,-----,~----,----,,----

-;fl 80
>-
A

(.)
z
60
LL
LL
w
..J 40
<t
>
0
e COD=TKN =6.07:t
20
A CQD:TKN=0.65=I
0
0
(.)
0 '----""-----'----------- .....___ ...___ ....___ _.
0 6 9 12 15 18 21 24
6c,days
Figure 3-a. Actual CODRemoval Efficiency Versus Mean Cell Residence
Time.

100
0
A

>-
(.) RESULTS ARE INDEPENDENT
z 80
OF COD=TKN RATIO
w
(.)
LL
LL
w 60
..J

0
w 40
a:::
0
0
(.) 20

3 6 9 12 15 18 21 24
6c,days
Figure 3-b. Theo~etical CODRemoval Efficiency Versus Mean Cell
Residence Time.
87

80
.....
OI
E 70
eC0D:TKN=6.07:I
z
A

0
•coD:TKN=0.65:1
t= 60
<1'.
0:
t- 50
z
w
(.)
z 40
0
(.)

30
0
0
(.)

t-
20
z
w
3
u.
10
u.
w 0
0 2 4 6 8 10 12 14 16 18 20 22
Be,days
Figure 4-a. Actual Soluble Effluent COD Concentration Versus Mean
..... Cell Residence Time
OI
E
z
A

70
0
t-
<1'.
0: 60
I-
z
w
(.) 50
z
0 RESULTS ARE INDEPENDENT
(.)
40 OF CQD:TKN RATIO
0
0
(.)
30
I-
z
w 20
::,
....I
LL
u. 10
w

0
0 2 4 6 8 10 12 14 16 18 20 22
(Jc,days
Figure 4-b. Theoretical Soluble Effluent COD Concentration Versus
Mean Cell Residence Time
88

4000---------------------------r---r--T---..--

3500 A

3000
......
[2500
z
2000
c::
1- 1500
z
w
(.)
z 1000 e C0D=TKN=G.07=1
0
(.) A C0D=TKN=0.65=1
en 500
en
....J 0 ..__...,_ _ _,__......,________________ ...__..._ ____ _.

0 2 4 6 8 10812 14 16 18 20 22
c,days
Figure 5-a. Actual Aeration Basin MLSS Concentration Versus Mean
Cell Residence Time.
4500 ----------------..------,----,-----,---,--

4000

3500

3000
......
e25oo
z
2000
<t
c::
1500
w
(.)
6 1000
(.)

en
CJ)
500
....J
0
0 2 4 6 8 10 12 14 16 18 20 22
Be,days
Figure 5-b. Theoretical Aeration Basin MLSSConcentration Versus
Mean Cell Residence Time.
89

100

90 •
80

0 70

_J
60
0
w 50
a:
z 40
I-
I- 30
z
w
(.)
20 e COD=TKN=6.07=1
a:
w •coD=TKN= 0.65=1
a..
10

00 2 4 6 8 10 12 14 16 18 20 22
8c,days

Figure 6. TKN Removal Efficiency Versus Mean Cell Residence Time.


90

100 ...---..---...----.--------------------

0
80
>-
(.)
z
w
u. 60
u.
w
...J

0 40
w e C0D=TKN= 6.07=1
a::
z AC0D=TKN=0.65=1
I 20
IO
:I:
z
0 ___ _.______ ___..__
__ _,______ ___,a _____ .,__ __ _,

0 3 6 9 12 15 18 21 24
Be,days
Figure 7-a. Actual NH3-N Removal Efficiency Versus Mean Cell
Residence Time.

';fl 80 \ ,. \_COD=TKN=6.07=1

>-
(.) \_COD=TKN=0.65 =I
z
w 60
LL
u.
UJ
...J
40
0
UJ
0::
z 20
I
ro
:I:
z
0
0 3 6 9 12 15 18 21 24
Be,days
Figure 7-b. Theoretical NH3-N Removal Efficiency Versus Mean Cell
Residence Time.
91

were higher for R-1 than for R-2 over the range of sludge ages

studied.

Percent Distribution of Effluent Nitrogen

Percent distribution diagrams of effluent nitrogen for the exper-

imental and theoretical study are presented in Figures 8-a and 8-b for

R-1 and in Figures 9-a and 9-b for R-2. As Sc increased, both the

experimental and theoretical data for R-1 indicated an increase in the

percentage of N03 -N whereas a decrease in the percentage of NH3-N

and N02 -N. Since TKN determinations were not conducted on the waste

sludge during the laboratory investigation, the amount of nitrogen

assumed to be incorporated into the biomass was calculated by sub-

tracting the summation of effluent TKN, NO;-N and NO;-N concentra-

tions from the influent TKN concentration. The experimental data

(Figure 8-a) indicated that nitrogen in the waste sludge decreased

while the theoretical data (Figure 8-b) indicated that nitrogen in the

waste sludge increased as the mean cell residence time increased.

The distribution diagram for the experimental data for R-2

(Figure 9-a) indicated that effluent TKN decreased as 8 was


C

increased, whereas N02 -N increased. Nitrate levels in the effluent

continued to increase with increases in mean cell residence time to

make up 8.5% of the influent TKN at a 8 = 1.5.7 days. A data point


C

fore = 16.5 days is not shown since nitrite data were not taken,
C

therefor8 the distribution diagram could not be completed. Approxi-

mately 1.0% of the influent TKNwas incorporated into the waste sludge
92

£100
w
r----,----,-----,r----r-----r-------~--~
(!)
0
a:
80
LL
0
z 60 Effluent TKN
0
I-
=>
m
• e Effluent 3
N0 -N
A Effluent N02-N
a: - N in Waste Sludge
I- 40
en
0

I-
z 20
w
0
a:
w
a.
0
0 3 6 9 12 15 18 21 24
IJc,days
Figure 8-a. Effect of Mean Cell Residence Time on Actual Percent
Distribution of Nitrogen for COD:TKN= 6.07:1.
OA
z
w
l!)
100
0
a: Effluent N03-N~
1-
z 80
ll..
0
z
0
1- 60
=>
m
a::
I-
C/) / Nin Sludge
c 40
1-
z / / Effluent N02-N
w
ua:: 20
w
a.

3 6 9 12 15 18 21 24
IJc,days
Figure 8-b. Effect of Mean Cell Residence Time on Theoretical Percent
Distribution of Nitrogen for COD:TKN= 6.07:1.
93

100
z
w
(!)
Effluent TKN
0 e Effluent N 0 3-N
a:: 80
!:::: A Effluent N0z-N
z -N in Waste Sludge
LL
0
60
iZ
0
I-
:::::,
CD
40
a::
I-
en
0
I- 20
z
w
(.)
a::
w
a.. 0
0 3 6 9 12 15 18 21 24
Bc,days
Figure 9-a. Effect of Mean Cell Residence Time on Actual Percent
Distribution of Nitrogen for COD:TKN= 0.65:1 (Nitrite
Oxidation Inhibited).
0 ~ 100
z
w
(!)

~----- Effluent N02,-N


t- 80
z
LL
0
z 60
Q
t-
:::::,
CD Effluent NH3 N
a:: 40
t- /Effluen1 NO--N
en 3
0
t- 20
z /Nin Waste Sludge
w
(.)
a::
w
a.. 0
0 3 6 9 12 15 18 21 24
8c,days
Figure 9-b. Effect of Mean Cell Residence Time on Theoretical
Percent Distribution of Nitrogen for COD:TKN= 0.65:1
(Nitrite Oxidation Inhibited).
94

ate = 2.2 days and increased to 5.5% ate = 8 days at which time it
C C

started decreasing to about 3.0% of the influent TKN at a sludge age

of 15.7 days. At a mean cell residence time of 8.0 days, both the TKN

and No;-N levels in the effluent tended to stabilize and each repre-

sented approximately 53% and 37% of the nitrogen leaving the

reactor. Both the NH3-N and No;-N levels decreased as ec increased,

while No;-N increased for the theoretical data (Figure 9-b). Nitrate

nitrogen decreased from 22% ate = 4.3 days down to 0% ate = 16.5
C C

days. At a mean cell residence time of 4.3 days, nitrite started to

increase reaching a maximum value of 97% at a sludge age of 16.5

days. Nitrogen that was incorporated into the waste sludge varied

from 4.0% of the total influent nitrogen at a sludge age of 2.2 days

to 2.5% of the total influent nitrogen at ec =16.5 days. NH3-N made

up only a small percentage of nitrogen leaving the reactor, varying

from 5.5% at ec = 2.2 days to 0.2% at ec = 16.5 days.

Destruction of Alkalinity

Figure 10 shows the actual amount of alkalinity destroyed in mg/1

as Caco 3 per mg/1 of TKN removed as a function of ec. For R-1, the

alkalinity consumed:TKN removed ratio increased as e increased reach-


c
ing a maximum value of 3.55 at a mean cell residence time of about 12

days. Alkalinity destroyed in mg/1 as Caco 3 per mg/1 of TKN removed

was much greater in R-2 over the range of sludge ages studied increas-

ing from 3.4 at ec = 2.2 days up to 5.55 at ec = 8.5 days at which it


0

"".0.5
0 e C0D 1 TKN = 6.07•1
w • C0D•TKN =0.65•1
>
0 -1.0
:::l!:
w
a:: -1.5
z
....
:::.:::
-2.0
.......
Cl
E
-2.5 •
.......
0
w -3.0
>- \0
0 V,

....
0::
Cl)
-3.5
w
0
-4.0
>-
!:::
..J
-4.5
<{

..J
<{
-5.0
rt)
0 -5.5
(.)
C
(.)

....... -6.0
Cl
E
-6.5
0 2 4 6 8 KJ 12 14 16 18 20 22
8 c,days
Figure 10. Alkalinity Destroyed mg/1 Caco3 /mg/1 TKNRemoved Versus Mean Cell Residence Time,
96

stabilized. Actual alkalinity change versus theoretical or predicted

alkalinity change is shown in Figure 11 using equation [4].

Biokinetic Constant Determination

The data collected on Reactors 1 and 2 were used to determine the

maximum yield coefficient (Ymax) and the energy of maintenance or

endogenous coefficient (kd) of the activated sludge for each

reactor. A plot of the net specific growth rate (1/e) versus the
C

specific substrate utilization rate (U) according to the following

equation is widely used in the environmental engineering literature

for determining Ymax and kd [122].

1
-=
max U - kd
Y [35]

Linearization of the experimental data collected on R-1 and R-2

according to Equation [35] is shown in Figure 12. The biokinetic

coefficients Ymax and kd were determined from a least squares linear

regression of the data to be 0.383 mg MLSS/mg CODand -0.057 days- 1

for R-1 and 0.318 mg MLSS/mg COD and+ 0.017 days-l for R-2.

OXYGENTRANSFERSTUDYRESULTS

In addition to the bench scale laboratory studies conducted in

this investigation, side studies involving oxygen transfer were also

initiated. Steady state and nonsteady state testing procedures were

used to determine oxygen transfer coefficients (K1 a) in order to eluc-

idate the relationship between the overall oxygen transfer coefficient


97

--
rt)

-
0
)(

0
rt)
4


u
COO:TKN =6.07:1


C
u 2
(/) COO:TKN = 0.65:1
<(

.......
0
C,
E
-2
>-
1--
z -4
::J
<(

-'
<(
-6

z
-8
w
(!)
z -10
<(
J:
u
Cl -12
w
0::
::::,
(/) -14
<(
w
-16
-16. -14 -12 -10 -8 -6 -4 -2 0 2 4
PREDICTED CHANGE IN ALKALINITY ,mg/I AS CaC0 3 {xt0 3 )

Figure 11. Relationship of Alkalinity Predicted by Equation [4]


to Measured Alkalinity Change.
0.50

045
A C0D=TKN=0.65=1
040
l =0.330U + 0.009
IJc
0.35

0.30
\0
0.25 00
U)
>-
C
"0 0.20
........

0.15
ft

0
........
0.10
e C0D=TKN=6.07=1
J_ =0.383U-0.057
IJc
0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 I.I 1.2 1.3 1.4 1.5
U, I/days

· Figure 12. Specific Growth Rate Versus Specific Substrate Utilization Rate.
99

and the oxygen uptake rate (R). Tables X and XI contain the oxygen

transfer data obtained on Reactors 1 and 2 during the course of the

experimental study.

Oxygen Transfer Coefficients (KL~

Steady state oxygen transfer coefficients (KLa) for R-1 increased

from 4.44 mg/1/hr ate


C
= 3.9 days to 6.42 mg/1/hr ate
C
= 21.0 days
except for the reading at 8.5 days which was 3.48 mg/1/hr. Oxygen

transfer coefficients determined on the wastewater mixed liquor by

steady state testing increased from 5.77 hr-lat a mean cell residence

time of 2.2 days to 26.58 hr- 1 at a ec value of 16.5 days.

Nonsteady state reaeration tests conducted on the effluent from

R-1 yielded KLa values that were virtually the same, averaging 26.69

hr- 1 (Table X). A listing of the oxygen transfer coefficients deter-

mined on the effluent from R-2 exhibited considerable variation. An

average nonsteady state KLa value of 43.40 hr-l was obtained during

the experimental study.

Oxygen Uptake Rates (R)

Figure 13 illustrates the effect of mean cell residence time and

the COD:TKNratio on the oxygen uptake rate of the mixed liquor sus-

pended solids. Results of the experimental investigation revealed

that oxygen uptake rate varied from 19.26 mg/1/hr at a mean cell resi-

dence time of 8.5 days to 41.24 mg/1/hr at a e of 13.8 days for


C
Table X

Summary of Oxygen Transfer Data for COD:TKN= 6.07:1

Value for Given eC , days

Parameter 3.9 4.2 8.5 13.8 21.0

Dissolved Oxygen
Mixed liquor (mg/1) 3.60 2.10 3.90 3.50 3.65

Dissolved Oxygen Saturation


*Mixed liquor (mg/1) 9.71 9.46 9.43 10.46 9.12
1--'
0
0
KLa Coefficient
Steady state (hr- 1) 4.44 4.55 3.48 5.93 6.42
Nonsteady state (hr- 1) 26.82 25.87 27.27 27.20 26.31

o2 uptake rate
Mixed liquor (mg/1/hr) 27 .12 33.52 19.26 41.24 35.10
Specific o2 uptake ra e
Mixed liquor (days-)
1 0.669 0.694 0.276 0.425 0.346
a coefficient
Effluent 1.014 0.978 1.031 1.028 0.995

6 coefficient
Effluent 1.126 1.097 1.094 1.213 1.058

*DOSATvalues obtained in nonsteady state reaeration tests


Table XI

Summary of Oxygen Transfer Data for COD:TKN= 0.65:1

Value for Given eC , days

Parameter 2.2 4.3 7.9 15.7 16.5

Dissolved oxygen
Mixed liquor (mg/1) 5.65 2.90 0.85 0.90 0.35

Dissolved oxygen saturation


*Mixed liquor (mg/1) 9.36 9.13 9.95 9.95 9.67 ....
....
0

KLa coefficient
Steady state (hr- 1) 5.77 16.15 16.43 17.14 26.58
Nonsteady state (hr- 1) 45.85 43.87 44.75 38.87 43.66
o2 uptake rate
Mixed liquor (mg/1/hr) 21.40 100.61 149.54 155.14 247. 71
Specific o2 uptake raie
Mixed liquor (days-) 1.072 2.276 1.416 1.154 1.658
a Coefficient
Effluent 1.733 1.659 1.692 1.470 1.651

8 Coefficient
Effluent 1.086 1.059 1.154 1.154 1.122
*DOsATvalues obtained in nonsteady state reaeration tests
102

280
e COD=TKN=6.07=1
260
COD= 334 mg/ I
240
8: 9.5hr
ACOD=TKN = 0.65=1

... 220
..c:
COD=322 mg/I
8=10.1 hr
' 200
'E
c,,

ft

180
....
w
<l'.
0::
w 160
:::ii::
....
<l'.
a. 140
:::>
z 120
w
<!)
>-
X 100
0

80

60

40
• •
20

0
0 2 4 6 8 10 12 14 16 18 20 22
8c ,days
Figure 13. Effect of Mean Cell Residence Time and COD:TKNRatio
on Oxygen Uptake Rate.
103

R-1. The specific oxygen uptake rate of the mixed liquor ranged from

0.276 days- 1 to 0.694 days- 1 •

Oxygen uptake rates of the mixed liquor in R-1 were 21.40,

100.61, 149.54 and 247.71 mg/1/hr at the corresponding mean cell resi-

dence times of 2.2, 4.3, 7.9, 15.7 and 16.5 days. The data on speci-

fic oxygen uptake rate of the mixed liquor exhibited no trend with

regards to sludge age.

Relationship Between Oxygen Uptake Rate and K~

Figure 14-a illustrates the effect of oxygen uptake rate on the

steady state oxygen transfer coefficient (KLa). Experimental data

collected on each reactor exhibited the same results that KLa

increases as oxygen uptake rate increases. A plot of the specific

oxygen utilizaiton rate versus KLa (Figure 14-b) yielded the same

results.

Alpha Coefficient (a)

A bench scale batch reactor was utilized for conducting nonsteady

state tests on the wastewater effluent from Reactors 1 and 2 and also

on Blacksburg, Virginia tap water. Results of the experimental data

collected indicate that the a coefficients were reasonably consistent

with average values of 1.009 and 1.641 for R-1 and R-2 respectively.
104

300 -----r------,r-------------------
e C0D=TKN=6.07=I
C0D=334mg/l
8 =9.5hr
250 • C0D=TKN=0.65= I
C0D=322mg/l
...
.c: 8=10.1 hr
.......
.......
Cl
E
200
w
I-
<(
0::
w
<(
I-
a.. 150
:::)

z
w
(!)
>-
X
0
100

50

0 '--'---------"-----'-----------'..__ __ __,
0 5 10 20 25 30

Figure 14-a. Relationship Between Oxygen Uptake Rate and Oxygen


Transfer Coefficient.
105

3.0 -----------------...--------
• COD=TKN =6.07= I
COD = 334 mg/L
C 9 = 9.5 hr
"C
......
_ 2.5 A COD: TKN = 0.65: I

w
.. COD= 322 mg/L

.... 8=10.1 hr
<t
a:
Z 2.0
0
....
<t
N
...J
..,_ 1.5
=>
z
w
(.!)
>-
X
0 1.0
u
1.1.
u
w
a.

••
en 0.5


0.0
0 5 10 20 25 30

Figure 14-b. Relationship Between Specific Oxygen Utilization Rate


and Oxygen Transfer Coefficient.
106

Beta Coefficient (e)

Thee coefficients were also determined on the effluent from each

reactor by nonsteady state tests. Average values determined

fore were 1.118 for R-1 and 1.115 for R-2.


VI. DISCUSSION

An analysis and discussion of the results derived from this study

will be presented in this Chapter. The major subdivisions of the

chapter are: Discussion of Experimental and Theoretical Results, in

which the laboratory and modeling data are discussed and Discussion of

Oxygen Transfer Study Results, in which the data collected in the

oxygen transfer laboratory studies are analyzed and discussed.

DISCUSSION OF EXPERIMENTAL
ANDTHEORETICALRESULTS

The data and results of the laboratory and modeling studies have

been presented. An analysis of the experimental results obtained from

the laboratory will be presented followed by a comparison and discus-

sion of the theoretical results derived from the computer modeling.

Evaluation of the data was accomplished by using the mathematical and

stoichiometrical relationships developed by Sherrard and coworkers [9,

117] for modeling the completely mixed activated sludge process.

Chemical Oxygen Demand

For the experimental study (Figure 3-a), chemical oxygen demand

removal efficiency remained virtually the same over the entire range

of mean cell residence times at a COD:TKNratio of 6.07:1. However,

at a COD:TKNratio of 0.65:1 the efficiency of removal was signifi-

cantly less at the lower mean cell residence times than at the pre-

vious COD:TKNratio of 6.07:1. Lower carbon removal at the lower

107
108

COD:TKNratio probably resulted in the heterotrophs being inhibited by

the high levels of free ammonia (approximately 33 mg/1 as N)

and/or NO;-N (average value of 162 mg/1 as N) that were present. A

nitrite accumulation occurred at the lower COD:TKNratio which pos-

sibly interfered with the COD determinations. COD analyses were cor-

rected for the nitrite interference by adding 10 mg sulfamic acid per

mg nitrite nitrogen according to Standard Methods [80]. The addition

of the sulfamic acid may have decreased the sensitivity of the COD

test resulting in higher CODvalues.

At the COD:TK.Nof 6.07:1, the COD removal efficiencies obtained

from the modeling data (Figure 3-b) were in good agreement with the

actual laboratory data collected. Analysis of the theoretical data

generated by the computer model indicated that COD removal efficiency

is independent of the COD:TKNratio at ratios of 6.07:1 and 0.65:1.

Theoretical COD removal efficiencies derived from the simulation study

of R-2 operating at a COD:TKNratio of 0.65:1 were higher than the

actual removal efficiencies obtained at the same ratio. The computer

model did not contain an inhibition function for free ammonia or nit-

rite, therefore carbon removal was higher for the theoretical situa-

tion and lower under actual process conditions.

For both the experimental and theoretical studies, effluent COD

concentrations decreased as the mean cell residence time increased at

each COD:TK.Nratio (Figures 4-a and 4-b). The modeling data in Figure

4-b indicate that effluent COD concentration was independent of the

COD:TK.Nratio at the values investigated in this study. For the


109

COD:TKNratio os 0.65:1, effluent CODconcentrations obtained from the

laboratory study were considerably higher than the experimental efflu-

ent CODvalues collected at the 6.07:1 ratio. These higher CODvalues

probably reflect the nitrite interference to the COD test. Effluent

COD concentrations from the modeling study (Figure 4-b) were lower

than the values obtained from the experimental study. If residual COD

values of 12.2 mg/1 and 30.6 mg/1 are subtracted from the experimental

data collected·at the COD:TKNratios of 6.07:1 and 0.65:1, the model

predicts the effluent CODconcentrations quite well. Refinement of

the model to account for substrate or product inhibition from free

ammonia and/or nltrite would probably result in closer agreement with

the experimental values.

Mixed Liquor Suspended Solids Concentration

The theoretical and experimental total microorganism concentra-

tion in the aeration basin are plotted as a function of the mean cell

residence time in Figures 5-a and 5-b.

As predicted by the biokinetic equations, the mixed liquor sus-

pended solids increased as the sludge age increased at each COD:TKN

ratio studied. It is evident from these plots that high nitrogenous

wastewaters (COD:TKN= 0.65:1) can support a higher concentration of

solids. Theoretically, the heterotrophic population in each of the

reactors should have been the same since the actual influent CODfor

each was approximately 330 mg/1. Howeyer, the decreased CODremovals

in R-2 as compared to R-1 probably reflect a smaller population of


110

heterotrophs which may have been inhibited by the high levels of nit-

rite and free ammonia. Mixed liquor suspended solids concentrations

in the aeration basins of R-1 and R-2 differed because of the size of

the nitrifier population in each. After a sludge age of 4 days, R-2

(COD:TKN = 0.65:1) can support a larger nitrifier population than R-1

(COD:TKN= 6.07:1) since there is much more nitrogen available for

nitrification. The model predictions for mixed liquor suspended

solids concentrations (Figure 5-b) were considerably greater than the

actual solids concentrations at each C0D:TKN ratio. At the C0D:TKN

ratio of 6.07:1, there was good agreement between the actual and the-

oretical results except at a e value of 21.0 days. The difference


C

between the two may be attributed to a low solids concentration

obtained from the experimental study which indicates that R-1 may not

have been completely at steady state conditions ate


C
= 21.0 days.
Mixed liquor suspended solids concentration at a C0D:TKN ratio of

0.65:1 were approximately 500 mg/1 lower than the values predicted by

the model at each mean cell residence time. This most likely can be

explained by the nitrite accumulation, high free ammonia levels, and a

possible oxygen limitation under which R-2 was operating during the

laboratory study, resulting in a lower population of nitrifiers con-

sisting primarily of Nitrosomonads.

TKN and NH3-N Removal and Efficiencies

From a comparison of Figures 6, 7-a and 7-b obtained on the two

influent wastewaters with COD:TKNratios of 6.07:1 and 0.65:1, several


111

observations of the experimental and theoretical data were made.

Decreasing the COD:TKNratio of the influent wastewater will signifi-

cantly lower the removal efficiencies for both TKN and NH3-N. At both

of the stoichiometric ratios tested, both the total Kjeldahl and

ammonia nitrogen removals are considerably lower at the low mean cell

residence times. The experimental data indicate that when the sludge

age falls below 8 days there is a significant decrease in both TKN and

NH3-N removal efficiencies, especially at the lower C0D:TKN ratio of

0.65:1. Increasing the mean cell residence time enhances nitrifica-

tion resulting in higher TKN and NH3-N removal efficiencies. At

high aC values, the nitrifier population becomes well established and

nitrification is enhanced. As a result, decreased levels of TKN and

NH3-N are found in the effluent from the system, whereas increased

concentrations of nitrate will occur. At low sludge ages, the nitri-

fier population is removed or "washed out" of the system. This

results in lower TKN and less NH3-N being removed from the wastewater

at low aC values. It would appear that lower total Kjeldahl and

ammonia nitrogen removals will occur at low C0D:TKN ratios which can

probably be attributed to the nitrifiers being inhibited by the high

concentrations of free ammonia and nitrite present [52-55].

In the computer modeling study, ammonia removal was much greater

at each C0D:TKN ratio than was observed in the experimental study.

There was better agreement between the theoretical analysis and exper-

imental data at a COD:TKNratio of 6.07:1 than at 0.65:1. Experi-

mental data collected on R-1 which operated at a C0D:TKN= 6.07:1


112

(Figure 7-a) exhibited variable results. Only four data points are

plotted for NH3-N removal at the 6.07:1 COD:TKN ratio because at

a aC value of 3.9 days, deamination of the organic nitrogen present in

the bacto-peptone caused a 22% increase in NH3-N resulting in a nega-

tive removal efficiency. The curve of best fit through the data

(Figure 7-a) does not match the TKN removal efficiency curve and prob-

ably resulted from experimental error in conducting the ammonia

analyses. A comparison of Figures 7-a and 7-b reveals that the model

predicted higher ammonia removal efficiencies than did the experi-

mental values obtained at a COD:TKN= 0.65:1 and that better NH3-N

removal was accomplished at a lower COD:TKN ratio, which is the oppo-

site of what the laboratory results indicated. Higher NH3-N removal

efficiencies were predicted by the model since substrate inhibition

was not developed into the model to account for the nitrite buildup

and high levels of free ammonia which may have inhibited the nitri-

fiers during the laboratory investigation. The model also predicted

that higher efficiencies of removal for ammonia nitrogen would be

accomplished by low COD:TKNratios. This anomaly can be explained by

the fact that the model assumes the total nitrogen concentration in

the influent consisted only of NH3-N, whereas in the laboratory data,

the total influent nitrogen concentration actually included some

organic nitrogen along with NH3-N. Therefore, more ammonia is avail-

able for utilization by the biomass and higher removal efficiencies

would occur. In practice though, as indicated by the experimental


113

results, higher ammonia removal efficiencies occur at high C0D:TKN

ratios.

The increased removal efficiencies for TKN and NH3-N that

occurred at the higher COD:TKNof 6.07:1 cannot be attributed to more

nitrogen being incorporated into the biomass. At a specified mean

cell residence time, effluent NH3 -N will be constant regardless of the

C0D:TKN ratio of the influent. Therefore, the efficiency of removal

will increase significantly at higher influent nitrogen concentrations

even though the amount of nitrogen incorporated into cellular biomass

remains constant at any C0D:TKN ratio since the influent nitrogen

levels are reduced by nitrification. Using percent removal does not

actually reflect the degree of treatment accomplished in the activated

sludge process and should be used cautiously since it is dependent

upon the influent nitrogen concentration.

Percent Distribution of Effluent Nitrogen

Figures 8-a, 8-b, 9-a, and 9-b are plots of the percent distribu-

tion of nitrogen in the effluent versus mean cell residence time for

C0D:TKN ratios of 6.07:1 and 0.65:1 obtained from the experimental and

theoretical investigations. The 100 percent nitrogen value represents

the influent TKN in the experimental study and the influent NH3-N in

the theoretical study. At steady state conditions, the influent

nitrogen concentration must equal the effluent nitrogen concentra-

tion. For the laboratory study, total nitrogen concentration in the

effluent consisted of TKN, No;-N, NO;-N and synthesized nitrogen in


114

the waste sludge. In the simulation study, NH3-N, N02 -N, N03 -N and

synthesized nitrogen in the waste sludge were the individual nitrogen

forms that made up the total nitrogen concentration in the effluent.

The amount of nitrogen assumed to be incorporated into the biomass was

determined through calculation as previously stated.

After examining the nitrogen distribution diagrams, the following

general statements can be made. Laboratory data for each C0D:TKN

stoichiometric ratio indicate that with increases in mean cell resi-

dence time, a decrease in TKN and an increase in nitrate levels was

observed in the effluent. At the C0D:TKNratio of 6.07:1, the amount

of nitrogen in the waste sludge increased slightly as the sludge age

increased, whereas the N02 -N levels decreased as ec increased. This

increase in synthesized nitrogen is attributed to an increase in the

nitrifier population as the sludge age increased. Under normal cir-

cumstances, nitrog~n in the biomass would increase slightly at the

onset of nitrification before decreasing as e increases, since less


C

biomass must be wasted from the system to maintain the desired mean

cell residence time. The increased levels of synthesized nitrogen at

the highest e value probably resulted from experimental error


C

associated with the ammonia testing and round-off error in calculating

the mass balances on nitrogen.

At the low COD:TKNratio of 0.65:1, the distribution of nitrogen

in the effluent is considerably different since R-2 was probably oper-

ating under a dissolved oxygen limitation at high mean cell residence

times. This resulted in a significant deterioration in the effluent


115

quality while operating under an oxygen restraint. Effluent N02 -N

levels were much greater than N03 -N levels indicating that nitrite

oxidation carried out by the Nitrobacter was severely inhibited.

Studies conducted on wastewater treatment systems treating high nitro-

genous wastes have experienced nitrite accumulations [SO, 52-55,

60-64] and the results obtained from this experimental study suggest

that the free ammonia concentration and the high concentration of

nitrite present inhibited both the Nitrosomonas and Nitrobacter

bacteria. It appears from an examination of Figure 9-a that

Nitrobacter was inhibited to a greater degree than Nitrosomonas.

Partial nitrification did occur at all mean cell residence times

greater than two days. Waste sludge nitrogen levels increased

slightly from approximately 3% of the influent TKN concentration

at ec = 2.2 days to about 5.5% at ec = 8 days before decreasing to

nearly 3% at a sludge age of 15.7 days. At low mean cell residence

times, there is no significant nitrifier propulation established;

therefore, synthesized nitrogen levels remain low and primarily result

from incorporation into the heterotrophic biomass. As 8 increases,


C

the nitrifier population increases incorporating more nitrogen into

the biomass. To increase the sludge age, less biomass must be wasted

from the system to maintain the desired 8 which results in a decrease


C

in nitrogen found in the waste sludge. Since partial nitrification

was occurring at sludge ages greater than 2.2 days, the oxidation of

ammonia helped to relieve the inhibition of the nitrifiers associated


116

with the high free ammonia concentration initially present at low ec

values.

The nitrogen distribution diagram developed from modeling R-1

(Figure 8-b) agrees well with the distribution diagram developed from

the lab data while the modeling of R-2 resulted in a significantly

different distribution of nitrogen (Figure 9-b) as compared to the

experimentally derived diagram. Predicted effluent nitrogen distribu-

tion at the COD:TKNratio of 6.07:1 was consistent with the actual

effluent nitrogen distribution except for a slight deviation in the

nitrogen in the waste sludge. The model predicts a decrease in waste

sludge nitrogen as e increases while the actual laboratory data show


C

nitrogen in the waste sludge to increase slightly. Differences

between the two diagrams result from experimental error and calculat-

ing the nitrogen in the waste sludge by difference rather than through

experimental analysis.

Results of the modeling study on R-2 operating at a COD:TKNratio

of 0.65:1 (Figure 9-b) indicated that nitrite accumulation occurred as

the mean cell residence time was increased. Oxygen transfer was

restricted to 56.7 millimoles which resulted in insufficient quan-

tities of oxygen being supplied to fulfill the requirements for com-

plete nitrite oxidation at long mean cell residence times. As a

result, nitrification was inhibited due to lack of oxygen and because

of the nitrite buildup which inhibited the nitrifiers. Synthesized

nitrogen decreased as the mean cell residence time increased. The

nitrate percentages predicted by the model were higher than the


117

experimental values. This may be attributed to nitrite interfering

with the Brucine Method [115] for determining nitrates which resulted

in low values being reported for nitrate values obtained in the

experimental study.

Destruction of Alkalinity

Figure 10 shows the effect of mean cell residence time and

COD:TKNratio upon the ratio of alkalinity destroyed in mg/1 as Caco 3

per mg/1 of TKN removed from the experimental and theoretical study.

Salient points inferred from this plot are: 1) mg/1 of alkalinity

destroyed per mg/1 of TKN removed increases as 6 increases and


C

2) mg/1 of alkalinity destroyed per mg/1 of TKN removed increases when

the COD:TKNratio of the influent wastewater is decreased.

As the mean cell residence time increases, nitrification effects

are enhanced, which result in hydrogen ions being liberated into solu-

tion, destroying the alkalinity. Nitrification increases at high eC


values because longer sludge ages approach complete oxidation of

ammonia nitrogen to nitrate nitrogen. Decreasing the COD:TKNratio or

increasing the ammonia concentration of the influent wastewater will

result in increased alkalinity destruction since more hydrogen ions

will be liberated when the ammonium ion (NH:) is oxidized to No;-N by

the nitrifiers.

A plot of the measured change in alkalinity in mg/1 as eaco 3

versus predicted change in alkalinity in mg/1 as Caco 3 (as calculated

from Equation [4]) is shown in Figure 11. The line of best fit
118

through the data points appears to be reasonably close to the theoret-

ical curve. Differences between the two may be attributed to experi-

mental error and problems associated with obtaining precise nitrogen

mass balances.

Biokinetic Constant Determination

Figure 12 is a plot of the net specific growth rate (1/e) versus


C

the specific substrate utilization rate (U) according to Equation

[35]. The biokinetic constants, Ymax and kd, determined from experi-

mental data collected on R-1 were considerably different from the

constants determined from the data collected on R-2.

A recent paper [123] has discussed some of the effects of nitri-

fication and COD:TKNratio on the determination of Ymax and kd.

Nitrification is an important factor that should not be overlooked in

developing biokinetic constants. Biokinetic coefficients developed in

activated sludge systems undergoing nitrification are significantly

different from those obtained under non-nitrifying conditions. Apply-

ing biokinetic constants that are representative of the total micro-

bial biomass rather than the individual coefficients for the hetero-

trophic and autotrophic reactions can lead to serious errors in the

model's predictability because of using invalid values for the maximum

yield coefficient (Ymax> and endogenous decay coefficient (kd).

Incorrect estimates of nutrient requirements, daily waste sludge pro-

duction, effluent substrate concentration, and oxygen requirements

will result if the proper, individual constants are not utilized in


119

the model. Increasing the ammonia level, i.e. decreasing the COD:TKN

ratio of the wastewater causes the specific growth rate versus

specific substrate utilization curve to shift to the left and results

in a different intercept and slope. This occurs since there is a

larger nitrifier population present at the lower COD:TKNratio result-

ing in a lower value for the specific substrate utilization rate at

the same mean cell residence time. The values for Ymax increase

whereas the values for kd tend to decrease.

The inclusion of data points obtained under non-nitrifying condi-

tions with data points obtained under nitrification for evaluation

biokinetic coefficients will severely alter the values for Ymax and

As a result, the curve for 1/e verus U is shifted to the right


C

yielding a milder slope and a more positive intercept. Therefore, the

value for Ymax decreases in magnitude while kd becomes more posi-

tive. A positive value for the endogenous coefficient has no mean-

ing. Only data point obtained under nitrifying conditions should be

used for evaluating Ymax and kd.

Originally, the coefficients determined from a linearization of

the data collected on Reactors 1 and 2 were averaged together for use

in the mathematical model. A better model prediction was achieved by

using the Ymax and kd coefficients from R-1 since they represented

typical values that have been reported in the literature. Since R-2

was designed to be oxygen limited at high aC values, inhibition of the

nitrifiers occurred due to the high concentration of nitrite and free

ammonia present in the mixed liquor. Since nitrification was


120

partially inhibited and the value of kd was positive for R-2, only the

biokinetic coefficients derived from the data collected on R-1 was

used in the simulation study.

DISCUSSIONOF OXYGENTRANSFERSTUDYRESULTS

In this section, an analysis and discussion of the data and

results obtained during the oxygen transfer laboratory studies will be

presented. A discussion of the oxygen transfer data collected on R-1

will be presented first followed by a comparison and analysis of the

data obtained on R-2. Tables X and XI contain the oxygen transfer

data collected on Reactors 1 and 2.

Oxygen Transfer Coefficients (Kcl

Both steady state and nonsteady state techniques were employed to

measure the overall oxygen transfer coefficient (KLa). Steady state

KLa values were obtained at each mean cell residence time by substi-

tuting the desired oxygen concentration of the mixed liquor, dissolved

oxygen saturation concentration of the mixed liquor (obtained through

nonsteady state reaeration tests) and the oxygen uptake rate of the

mixed liquor into Equation [18]. Tables X and XI show that the steady

state KLa values increased as the sludge age increased for both R-1

and R-2, except at a ec value of 8.5 days the R-1 steady state KLa

value of 3.48 hr- 1 was less than the value of 4.55 hr- 1 ate
C
= 4.2
days. The lower KLa value resulted since the oxygen uptake rate was

considerably lower at this sludge age than at the previous mean cell
121

residence time. Steady state oxygen transfer coefficients for R-2

were considerably greater than those for R-1, especially at the longer

mean cell residence times. These larger KLa values probably can be

attributed to the low dissolved oxygen concentrations at the higher

sludge ages resulting in a greater driving force for oxygen trans-

fer. This observation~ that KLa increases as the sludge age increases

signifies that the oxygen transfer coefficient for a particular aera-

tion system is not constant but variable which agrees with the find-

ings of Albertson and DiGregorio [113] and Eckenfelder [114].

Separate nonsteady state reaeration tests were conducted on the

effluent from each reactor system to determine the nonsteady state

oxygen transfer coefficients. Analysis of the tabulated nonsteady

state KLa values in Tables X and XI indicated that there is no trend

with the mean cell residence time for either reactor. The KLa coeffi-

cients determined on the effluent from R-2 were approximately 1.5

times the KLa coefficients determined for R-1. Nonsteady state test-

ing procedures indicate that KLa is a constant and does not vary with

sludge age. According to standard aeration theory, KLa is a constant

for a given aerator.

The discrepancy between the two testing methods for determining

KLa is rather difficult to explain. It must be remembered that the

steady state tests were conducted in mixed liquor, therefore oxygen

transfer into the system is affected by a biological reaction, the

oxygen uptake rate of the mixed liquor suspended solids. In the non-

steady state tests, wastewater effluent was utilized and involves only
122

a physical mass transfer of oxygen and exclud~s any biological reac-

tions that may be occurring under steady state conditions in the aera-

tion basin. Oxygen transfer in non-Newtonian fluids such as activated

sludge and fermentation broths is quite different from that in

Newtonian liquids [22]. A third explanation suggests that the pres-

ence of cells affects the oxygen transfer rates usually resulting in a

higher value than that of the physical transfer rate obtained from

clean water testing [22, 86, 113]. Finally, the standard BOD bottle

testing procedure [80] may be an invalid technique for determining the

oxygen uptake rate of the mixed liquor suspended solids.

Oxygen Uptake Rates (R)

The effect of mean cell residence time and the COD:TKNratio on

the oxygen uptake rate of the mixed liquor suspended soldis (R) was

presented in Figure 13. For both Reactors 1 and 2, oxygen uptake

increased as the mean cell residence time increased. This is expected

since higher sludge ages enhance nitification which requires more

oxygen. For R-2 operating at a COD:TKN= 0.65:1, the oxygen uptake

rate was much greater than that of R-1 which operated at a COD:TKN

ratio of 6.07:1. The theoretical total oxygen demand of the influent

wastewater to R-2 was approximately five times that of R-1 and theo-

retically, the same amount of air (1 liter/minute) was being supplied

to both reactors. The average dissolved oxygen concentration in the

mixed liquor was approximately 3.35 mg/1 during the laboratory studies

for R-1. This value is 9.5 times the DO level in R-2 at a mean cell
123

residence time of 16.5 days indicating that R~2 was probably oxygen

limited at high mean cell residence times and exhibited a much greater

oxygen uptake rate than that of R-1 because of enhancement of nitri-

fication.

Relationship Between Oxygen Upate Rate and K~

Figure 14 illustrates the effect of oxygen uptake rate on the

steady state oxygen transfer coefficient (KLa). It is evident from

this plot that KLa is directly related to the oxygen uptake rate of

the mixed liquor suspended solids. As the oxygen uptake rate

increased, the KLa of the aeration system increased for both R-1 and

R-2. Again, this is more evidence that KLa is not a constant for a

specified aeration device and agrees with previous findings [113,

114]. Figure 15 is a reproduction of the data collected by Albertson

and DiGregorio [113] illustrating the increase in the "corrected to

standard conditions" KLa with increases in the oxygen uptake rate of

the mixed liquor. They concluded that the direct path of oxygen

transfer from the gas bubbles to the suspended microorganisms was

responsible for the observed phenomenon. Other investigators have

also considered this direct oxygen transfer mechanism [86, 103,

111-113]. An analysis of the nonsteady state and steady state oxygen

transfer data collected by Kayser [95] as plotted in Figure 16 illus-

trates similar results, i.e., that KLa is not a constant.

The effect of specific oxygen utilization rate (SOUR) on the

steady state KLa value is plotted in Figure 14-b. Plotting the


124

16 0 -----.-----,.---T"----,-----,i----~

140

120

...
.c
' 100
'C,
E
w Surface aeration
80
a::
w
60
a..
:::)

z
w
0
>- 40
X
0

20
Submerged aeration

00 5 10 15 20 25 30

Figure 15. Oxygen Uptake Rate Versus ~a Corrected to Standard


Conditions (After Albertson and DiGregorio [113]).
50
• Nonsteady State
45
Instable Steady State
40
0:::
w
.. 35 •
0::: 30
w
25
t:!
a.. .....
N
:::) 20 V1

z
w 15
(.!)
>-
X
0 10

0

0 2 3 4 5 6 7 8 9 10 II 12 13 14 15 16
I
{Kla),1/hr

Figure 16. Oxrgen Uptake Rate as a Function of ~a (After Kayser [95]).


126

specific oxygen utilization rate eliminates the effects of the mixed

liquor suspended solids concentration on the oxygen uptake rate and is

a better parameter for comparative purposes. There is more

variability of the data plotted in Figure 14-b than in Figure 14-a.

Experimental errors in weighing the suspended solids can result in

considerably different values for SOUR. The line of best fit through

the data indicates that the steady state KLa values increases as a

function of the specific oxygen utilization rate.

Alpha Coefficient (a)

Effluent samples from each reactor were used in nonsteady state

tests for determining the alpha coefficients (a). Tables X and XI

show that the a coefficients for each wastewater were fairly consis-

tent during the course of the study. Alpha coefficients are utilized

to calculate the oxygen transfer coefficient under process conditions

when the oxygen transfer coefficient is known for a specified aeration

device tested in tap water at zero dissolved oxygen concentration and

one atmosphere of pressure. For the effluent from R-2, the a coeffi-

cients are approximately 50% greater than those for R-1. Effluent

characteristics from R-2 were such that the overall oxygen transfer

coefficients were nearly 1.5 times the KLa values determined in the

effluent from R-1. As a result, the alpha values are considerably

greater than those for R-1. The KLa coefficient of the tap water used

in the nonsteady state tests for calculating the alpha values was

determined to be 26.45 hr-lat 20°c.


127

The increased transfer rates observed in the effluent from R-2

probably resulted from smaller bubbles that were observed during the

nonsteady state reaeration tests. Larger sized bubbles were observed

during nonsteady state testing of the effluent from R-1 resulting in

lower values for KLa. A higher surfactant concentration possibly

occurred in R-2 since the process was oxygen limited and lower carbon

removals were observed. The presence of minute concentrations of

surfactants can result in smaller sized bubbles developing which

enhance oxygen transfer. Barnhart [124] states that the bubble size

decreases when surfactants are present which result in an increase in

the oxygen transfer coefficient (KLa) and alpha coefficient.

Beta Coefficient (e)

The beta coefficient which is the ratio of the dissolved oxygen

saturation concentration of the wastewater to the dissolved oxygen

saturation concentration of the tap water was determined during the

nonsteady state testing. Tables X and XI contain thee coefficients

for each of the wastewaters studied. The coefficients show some var-

iability and do not indicate a trend with the mean cell residence

time. During the experimental study, the average value fore was

1.118 and 1.115 for R-1 and R-2 respectively. Normally, e values

less than one would be expected in wastewater. Thee coefficients in

this study may have been greater than one because the wastewater may

have become slightly supersaturated with oxygen since a mixer and

diffuser stone were used for transfering air into the reactor. This
128

probably resulted in more oxygen being entrained in the wastewater due

to the characteristics of the synthetic wastewater as compared to tap

water.

SUMMARY

The results and discussion of the laboratory studies, theoretical

modeling simulations and oxygen transfer studies have been pre-

sented. A brief summary of the salient points of this research inves-

tigation will now be presented.

Higher organic carbon removal efficiencies were obtained from

experimental data on R-2 operating at a COD:TKNratio of 6.07:1.

Lower carbon removal at a COD:TKNratio of 0.65:1 probably resulted

from inhibition as a result of the high concentration of free ammonia

and nitrite present in the mixed liquor. R-2 was theoretically

designed to operate under a carbon limitation at low mean cell resi-

dence times and under a dissolved oxygen limitation at high sludge

ages. The low DO levels and high oxygen uptake rates suggest that R-2

was indeed operating under a dissolved oxygen limitation at high

sludge ages during the laboratory study. Nitrification can be inhi-

bited when insufficient amounts of oxygen are supplied to the process

and result in a nitrite buildup [35, 50, 54]. The theoretical model

predictions for carbon removal were found to be independent of the

COD:TKNratios studied. CODremoval efficiency data generated on R-2

exhibited the greater difference between theoretical and actual pro-

cess conditions. Refinement of the model through addition of inhibi-


129

tion functions for nitrite and free ammonia should result in better

predictability of the model.

Total mixed liquor suspended solids concentrations increased for

both the experimental and modeling results as sludge age was

increased. A higher microorganism concentration existed in R-2

because of the larger nitrifier population which could be supported by

the high levels of NH3 -N available for nitrification. Model predic-

tions were greater than actual microbial concentrations with the larg-

est difference found in R-2. Use of biokinetic constants selected

from the literature contributed to part of this difference in agree-

ment. A second reason would suggest that the heterotrophic population

was smaller than the predicted concentration because the high concen-

tration of nitrite and free ammonia probably inhibited the hetero-

trophs resulting in the low carbon removals exhibited for R-2 during

the laboratory studies.

TKN and NH3 removal efficiencies were greater at the higher

COD:TKNratio of 6.07:1 and as the sludge age increased for the exper-

imental studies. Results of the modeling study indicated that ammonia

removal was greater than what was achieved in the experimental runs

and that better removal could be accomplished at the lower COD:TKN

ratio of 0.65:1. This anomaly results from an assumption in the model

that considered influent nitrogen to consist only of NH3-N when in

fact the TKN of the actual waste contained both organic and ammonia

nitrogen.
130

Experimental and theoretical data for percent distribution of

effluent nitrogen were considerably different for R-1 and R-2. Model

predictions for R-1 were much closer in agreement with actual values

since R-2 was inhibited by the presence of high free ammonia and

nitrite levels. The actual and theoretical data for R-1 indicated

that a nitrate buildup would occur whereas a nitrite accumulation

would result in R-2. A DO limitation probably existed in R-2 at

high eC values resulting in Nitrobacter being inhibited and therefore

causing a nitrite buildup. A better agreement between theory and

process conditions could be obtained by refining the model to account

for substrate and produce inhibition along with the use of actual

biokinetic constants rather than literature values for the autotrophic

reactions.

The amount of alkalinity destroyed in mg/1 as eaco3 per mg/1 of

TKN removed increased as sludge age increased and was greater at the

COD:TKNratio of 0.65:1. More NH3-N was available for nitrification

in R-2 therefore, a greater potential for alkalinity destruction

existed. Theoretical and experimental changes in alkalinity were

found to be in close agreement. If nitification had not been inhi-

bited in R-2 and better nitrogen balances could have been calculated

from the laboratory analyses, a better match between practice and

theory should have resulted.

Total system biokinetic constants (Ymax and kd) were determined

by plotting net specific growth rate (1/e) versus specific substrate


C

utilization rate (U). Decreasing the COD:TKNratio of the wastewater


131

will cause the above curve to shift upward and to the left resulting

in a larger value for Ymax and a more positive value for kd. Includ-

ing data points obtained under non-nitrifying conditions with data

points obtained under nitrifying conditions can severely alter the

results of ~ax and kd. Under nitrifying conditions and low COD:TKN

ratios, a larger nitrifier population exists which results in a lower

value for the specific substrate utilization rate. The values for

Ymax increases whereas the value for kd tend to decrease.

Evaluation of the oxygen transfer data revealed several interest-

ing results. Steady state oxygen transfer coefficients increased as

mean cell residence time increased at each COD:TKNratio studied. K1 a

values for R-2 were considerably greater than those for R-1 since R-2

was probably operating under a DO limitation at high ec values.

Nonsteady state K1 a values were essentially constant as a function of

sludge age. Oyxgen transfer coefficients obtained under nonsteady

state conditions for R-2 were approximately 1.5 times greater than the

values for R-1. Explanation of this finding can be attributed to the

smaller bubbles observed in the effluent from R-2 during the nonsteady

state tests which probably resulted from a high surfactant concentra-

tion that could have existed due to low organic removals which

resulted in higher organic concentrations in the mixed liquor of

R-2.

Oxygen uptake rates of the mixed liquor generally increased

as e increased for both reactors with a more pronounced increasQ for


C

R-2. Mixed liquor DO levels for R-2 were considerably lower than
132

those for R-1 at high sludge ages indicating a possible oxygen limita-

tion existed.

The most significant observation of the study was that the steady

state KLa values were found to increase as the oxygen uptake rate

increased. This observation indicates that KLa is not a constant

value but a variable and that it is dependent on the microbial respir-

ation rate. Conventional aerator design and operation will have to be

reevaluated because of the impact of this finding.

Both the alpha and beta coefficients determined on the effluent

from each reactor were fairly consistent. There were no trends indi-

cated for either as a function of mean cell residence time.


VII. CONCLUSIONS

Two continuous flow completely mixed activated sludge units were

operated in the laboratory for approximately nine months. Biokinetic

and stoichiometric equations were used to evaluate the data and a

mathematical model was developed to simulate the process. Oxygen

transfer studies were conducted on the mixed liquor and effluent from

the reactors. From an analysis of the data collected and generated

during this investigation, the following conclusions are made:

1. The oxygen transfer coefficient (KLa) determined from steady

state testing generally increased in magnitude as the mean cell resi-

dence time increased. Determination of oxygen transfer coefficients

from nonsteady state testing indicated that KLa was constant for dif-

fused aeration and dependent on the characteristics of the wastewater

utilized in the analysis.

2. The oxygen transfer coefficient of the system was found to

increase as both the oxygen uptake rate of the mixed liquor suspended

solids and specific oxygen utilization rate increased, indicating that

KLa is not a constant for a given aeration device. This can most

likely be attributed to direct oxygen transfer from the bubble to the

microorganism. A second explanation suggests that the standard BOD

bottle testing procedure for determining oxygen uptake rate may be

invalid.

3. Operation of the completely mixed activated sludge process

approaching a dual substrate limitation where carbon is limiting at

133
134

low mean cell residence times and oxygen limiting at high mean cell

residence times resulted in a deterioration in effuent quality at high

mean cell residence times. Nitrification was observed to be inhibited

and nitrite nitrogen and ammonia nitrogen were found to accumulate in

the effluent.

4. The use of a mathematical model employing biokinetic and

stoichiometric equations can be used for predicting effluent quality

from the activated sludge process when operation approaches a dual

substrate limitation.

5. The alpha (a) and beta (S) coefficients determined from non-

steady state testing showed no relationship to mean cell residence

time.
VIII. RECOMMENDATIONS
FOR FUTURESTUDY

Examination of th~ results of this research indicates further study

is warranted. The following topics are suggested for further investiga-

tion:

1. Operation of the activated sludge process under dual substrate

limitations with carbon and phosphorus being the limiting

nutrients.

2. Evaluation of the BOD Bottle Testing Procedure for determining

the oxygen uptake rate of the mixed liquor suspended solids.

3. Investigation of nitrite interference with the Brucine Method

for determining nitrate nitrogen.

4. Evaluation of the settling and dewatering characteristics of

the sludge produced under dual substrate limitations.

5. Examination of nitrite interference on the determination of

CODemploying the Dichromate Reflux Method and sulfamic

acid.

135
REFERENCES

1. Ardern, E. and Lockett, W. T., "Experiments on the Oxidation of


Sewage Without the Aid of Filters." Journal Society
Chemical Industry, 33, 523-539, 1122-1124 (1914).

2. Lawrence, A. W. and McCarty, P. L., "Unified Basis for Biological


Treatment Design and Operation." Journal Sanitary
Engineering Division, ASCE, ~' 757-778 (1970).

3. McKinney, R. E., "Mathematics of Complete-Mixing Activated


Sludge." Journal Sanitary Engineering Division, ASCE, ~'
87-113 (1962). --

4. Goodman, B. L. and Englande, A. J., "A Unified Model of the


Activated Sludge Process." Journal Water Pollution Control
Federation, 46, 312-331 (1974).

5. Adams, C. E. and Eckenfelder, W.W., "A Kinetic Model for Design


of Completely-Mixed Activated Sludge Treating Variable-
Strength Industrial Wastewaters." Water Research, -2._,37-42
(1975).

6. Sherrard, J. H., "Kinetics and Stoichiometry of Completely Mixed


Activated Sludge." Journal Water Pollution Control
Federation,~' 1968-1975 (1977).

7. Stall, T. R. and Sherrard, J. H., "Evaluation of Control


Parameters for the Activated Sludge Process." Journal Water
Pollution Control Federation, 22._, 450-457 (1978).

8. Benefield, L. D. and Randall, C. W., "Evaluation of a


Comprehensive Kinetic Model for the Activated Sludge
Process." Journal Water Pollution Control Federation, 49,
1636-1641 (1977).

9. Sherrard, J. H. and Schroeder, E. D., "Stoichiometry of


Industrial Biological Wastewater Treatment." Journal Water
Pollution Control Federation,~' 742-747 (1976).

10. Sherrard, J. H., "Activated Sludge Wastewater Treatment


Stoichiometric Relationships." Journal Chemical Technology
and Biotechnology, 30, 447-452 (1980).

11. Hawkins, J.M. and Sherrard, J. H., "Stoichiometry of Phosphorus


Removal in the Activated Sludge Process." Civil Engineering
for Practicing and Design Engineers, l_, 91-106 (1982).

136
137

12. Hart, G. M. and Sherrard, J. H., "Characteristics of the


Activated Sludge Process Under Nitrogen Limiting
Conditions." in Proceedings: 15th Mid-Atlantic Waste
Conference, pg. 24,(1983).

13. Mines, R. O. and Sherrard, J. H., "Effect of Oxygen Requirements


on the Stoichiometry of Activated Sludge." Proceedings
International Symposium on Gas Transfer at Water Surfaces,
Cornell University, Ithaca, New York, June 13-15, 1983 (in
press).

14. Novak, J. T., Mines, R. O. and Sherrard, J. H., "Activated Sludge


Processes and Effluent Standards." Journal Water Pollution
Control Federation, 55, 332-335 (1983).

15. Monod, J., "The Growth of Bacterial Cultures." Annual Review of


Microbiology, Volume 3, 371-394 (1949).

16. Herbert, D., Elsworth, R. and Telling, R. C., "The Continuous


Culture of Bacteria; A Theoretical and Experimental
Study." Journal of General Microbiology, 14, 601-622
(1956).

17. Cooney, C. L., Wang, D. I. C., and Mateles, R. I., "Growth of


Enterobacter aerogenes in a Chemostat with Double Nutrient
Limitations." Applied and Environmental Microbiology, 31,
91-98 (1976). -

18. Sykes, R. M., "Identification of the Limiting Nutrient and


Specific Growth Rate." Journal Water Pollution Control
Federation, 45, 888-895 (1973).

19. Sykes, R. M., "Theory of Multiple Limiting Nutrients." Journal


Water Pollution Control Federation,~, 2387-2392 (1974).

20. Ryder, D. N. and· Sinclair, C. G., "Model for the Growth of


Aerobic Microorganisms Under Oxygen Limiting Conditions."
Biotechnology and Bioengineering, 14 787-798 (1972).

21. Fredrickson, A.G., Megee, R. D., and Tsuchiya, H. M.,


"Mathematical Models for Fermentation Processes." Advances
in Applied Microbiology, 13, 419-465 (1970).

22. Sinclair, C. G. and Ryder, D. N., "Models for the Continuous


Culture of Microorganisms Under Both Oxygen and Carbon
Limiting Conditions." Biotechnology and Bioengineering, _!Z_,
375-398 (1975).
138

23. Herbert, D., in Recent Progress in Microbiology, VII


International Congress for Microbiology, ed. by G. Tunevall,
Almquist, and Wiksell, Stockholm (1958).

24. Bader, F. G., Meyer, J. S., Fredrickson, A.G. and Tsuchiya,


H. M., "Comments on Microbial Growth Rate." Biotechnology
and Bioengineering, 17, .279-283 (1975).

25. Megee, R. D., Drake, J. F., Fredrickson, A.G. and Tsuchiya,


H. M., "Studies in Intermicrobial Symbiosis. Saccharomyces
cerevisiae and Lactobacillus casei." Canadian Journal of
Microbiology, .!!!_, 1733-1742 (1972).

26. Cooney, C. L. and Wang, D. I. C., "Transient Response of


Enterobacter aerogenes Under a Dual Nutrient Limitation in a
Chemostat." Biotechnology and Bioengineering, 18, 189-198
(1976).

27. Bader, F. G., "Analysis of Double-Substrate Limited Growth."


Biotechnology and Bioengineering, 20, 183-202 (1978).

28. Randall, C. W., Marshall, D. W. and King, P.H., "Phosphate


Release in Activated Sludge Process." Journal Sanitary
Engineering Division, ASCE, 96, 395-408 (1970).

29. Shindala, A., "Nitrogen and Phosphorus Removal from


Wastewater - Part I." Water and Sewage Works, 119, 66-71
(1972).

30. Grady, C. P. L. and Lim, H. C., Biological Wastewater


Treatment. Marcel Dekker, Inc., New York, New York
(1980)

31. Stenstrom, M. K. and Poduska, R. A., "The Effect of Dissolved


Oxygen Concentration on Nitrification." Water Research, 14,
643-649 (1980).

32. Painter, H. A., "A Review of Literature of Inorganic Nitrogen


Metabolism in Microorganisms." Water Research, !!._,393-450
(1970).

33. Focht, D. D. and Chang, A. c., "Nitrification and Denitrification


Processes Related to Waste Water Treatment." Advances in
Applied Microbiology, .!2_, 153-186 (1975).

34. Focht, D. D. and Verstraete, W., "Biochemical Ecology of


Nitrification and Denitrification." Advances in Microbial
Ecology, _!_, 135-214 (1977).
139

35. Belser, L. W., "Population Ecology of Nitrifying Bacteria."


Annual Review of Microbiology, 33, 309-333 (1979).

36. Benefield, L.D. and Randall, C. W., Biological Process Design for
Wastewater Treatment. Prentice Hall, Inc., Englewood
Cliffs, New Jersey (1980).

37. Sharma, B. and Ahlert, R. C., "Nitrification and Nitrogen


Removal." Water Research, _!_!_,897-925 (1977).

38. Eckenfelder, W.W., Principles of Water Quality Management. CBI


Publishing Company, Inc., Boston, Massachusetts (1980).

39. United States Environmental Protection Agency, Process Design


Manual for Nitrogen Control. Office of Technology Transfer,
Washington, D.C. (1975).

40. Adams, C. E. and Eckenfelder, W.W., "Nitrification Design


Approach for High Strength Ammonia Wastewaters." Journal
Water Pollution Control Federation,~, 413-421 (1977).

41. Wuhrman, K., "Effects of Oxygen Tension on Biochemical Reaction


in Sewage Purification Plants." in Biological Waste
Treatment Processes, 3rd Conference on Biological Waste
Treatment, W.W. Eckenfelder and J. McCabe, eds., Pergamon
Press, New York, New York (1963).

42. Bennett, G. F. "Oxygen Mass Transfer Rates, Mechanisms, and


Applications in Biological Waste Water Treatment." CRC
Critical Reviews in Environmental Control, _2_, 300-39Z-
(1980).

43. Charley, R. C., Hooper, D. G. and McLee, A.G., "Nitrification


Kinetics in Activated Sludge at Various Temperatures and
Dissolved Oxygen Concentrations." Water Reseach, 14,
1387-1396 (1980). -

44. Wezernak, C. T. and Gannon, J. J., "Oxygen-Nitrogen Relationships


in Autotrohpic Nitrification." Applied Microbiology, 15,
1211-1215 (1967). -

45. Haug, R. T. and.McCarty, P. L., "Nitrification with Submerged


Filters." Journal Water Pollution Control Federation, 44,
2086-2102 (1972).

46. Kiesow, L.A., "A Biologic Assimilation of Inorganic Energy."


Aldrichimica Acta, 2._, 33-37, (1972).
140

47. Randtke, S. J. and McCarty, P. L., "Variations in Nitrogen and


Organics in Wastewaters." Journal of the Environmental
Engineering Di vision, ASCE, 103, 539-550 (1977) •

48. Wu, Y. C., "Role of Nitrogen in Activated Sludge Process."


Journal of the Environmental Engineering Division, ASCE,
102, 897-907 (1976).

49. Sherrard, J. H., "Determination of Activated Sludge Process


Oxygen Requirements." Civil Engineering for Practicing and
Design Engineers, _!_, 33-44 (1982).

SO. Laudelout, H., Lambert, R. and Pham, M. L., "Influent Du pH Et De


La Pression Partielle D'Oxygene Sur La Nitrification."
Annales De Microbiologie, 127A, 367-382 (1976).

51. Srinath, E.G., Loehr, R. c., and Prakasam, T. B. S., "Nitrifying


Organism Concentration and Activity." Journal Environmental
Engineering Division, ASCE, 102, 449-463 (1976).

52. Prakasam, T. B. S., Joo, Y. D., Srinath, E.G. and Loehr, R. c.,
"Nitrogen Removal from a Concentrated Waste by Nitrification
and Denitrification." Proceedings 29th Purdue Industrial
Waste Conference, Purdue University, W. Lafayette, Indiana,
Extension Series 145: 497-509 (1974).

53. Prakasam, T. B. S. and Loehr, R. C., "Microbial Nitrification and


Denitrification in Concentrated Wastes." Water Research, 6
859-869 (1972).

54. Anthonisen, A. C., Loehr, R. C., Prakasam, T. B. S. and Srinath,


E.G., "Inhibition of Nitrification by Ammonia and Nitrous
Acid." Journal Water Pollution Control Federation, 48,
835-852 (1976).

55. Verstraete, W., Vanstaen, H. and Voets, J.P., "Adaptation to


Nitrification of Activated Sludge Systems Treating Highly
Nitrogenous Waters." Journal Water Pollution Control
Federation, 49, 1604-1608 (1977).

56. Sherrard, J. H., Scearce, S. N., Bennin,¥.er, R. W. and Weber,


A. S. "Influence of the COD:TKN:NH4-N Ratio on Nitrification
and Alkalinity in the Activated Sludge Process." Journal
Civil Engineering Design, 2(1), 31-44 (1980).

57. Stankewich, M. J., "Biological Nitrification with the High Purity


Oxygenation Process." Proceedings 27th Purdue Industrial
Waste Conference, Purdue University, W. Lafayette, Indiana,
Extension Series 141: 1-23 (1972).
141

58. Hockenbury, M. R., Daigger, G. T., and Grady, C. P. L., "Factors


Affecting Nitrification." Journal Environmental Engineering
Division, ASCE, 103, 9-19 (1977).

59. Meiklejohn, J., "Some Aspects of the Physiology of the Nitrifying


Bacteria." in Autotrophic Microorganisms, 4th Symposium
Society General Microbiology, B. A. Fry and J. L. Peel,
eds., University Press, Cambridge.

60. Boon, B. and Laudelout, H., "Kinetics of Nitrite Oxidation by


Nitrobacter windgrad-skyi." Biochemical Journal, ~' 440-447
(1962).

61. Voets, J.P., Vanstaten, H. and Verstraete, W., "Removal of


Nitrogen from Highly Nitrogenous Wastewaters." Journal
Water Pollution Control Federation,~' 394-398 (1975).

62. Tomlinson, T. G., "Inhibition of Nitrification in the Activated


Sludge Process of Sewage Disposal." Journal Applied
Bacteriology, 22._, 266-291 (1966).
63. Murray, I., Parsons, J. W. and Robinson, K., "Inter-Relationships
Between Nitrogen Balance, pH and Dissolved Oxygen in an
Oxidation Ditch Treating Farm Animal Waste." Water Research,
.2_, 25-30 (197 5).

64. Hutton, W. C. and LaRocca, S. A., "Biological Treatment of


Concentrated Ammonia Wastewaters." Journal Water Pollution
Control Federation,!!]_, 989-997 (1975).

65. Keenan, J. D., Steiner, R. L. and Fungaroli, A. A., "Substrate


Inhibition of Nitrification." Journal Environmental Science
and Health, Al4(5), 377-397 (1979).

66. Wild, H. E., Sawyer, C. N. and McMahon, T. C., "Factors Affecting


Nitrification Kinetics." Journal Water Pollution Control
Federation, ~' 1845-1854 (1971).

67. Kholdebarin, B. and Oertli, J. J., "Effect of pH and Ammonia on


the Rate of Nitrification of Surface Water." Journal Water
Pollution Control Federation, 49, 1688-1692 (1977).

68. Poduska, R. A. and Andrews, J. F., "Dynamics of Nitrification in


the Activated Sludge Process." Proceedings 29th Purdue
Industrial Waste Conference, Purdue University,
W. Lafayette, Indiana, Extension Series 145: 1005-1025
(1974).
142

69. Knowles, G., Downing, A. L. and Barrett, M. J., "Determination of


Kinetic Constants for Nitrifying Bacteria in Mixed Cultures,
with the Aid of an Electronic Computer." Journal General
Microbiology, 38, 263-278 (1965).

70. Wong-Chong, G. M. and Loehr, R. C., "The Kinetics of Microbial


Nitrification." Water Research, 1_, 1099-1106 (1975).

71. La Motta, E. J. and Shieh, W. K., "Diffusion and Reaction in


Biological Nitrification." Journal Environmental
Engineering Division, ASCE, 105, 655-673 (1979).

72. Bass, S. J. and Shell, G. L., "Evaluation of Oxygen Transfer


Coefficients of Complex Wastewaters." Proceedings 32nd
Purdue Industrial Waste Conference, Purdue University,
W. Lafayette, Indiana, Vol. 32, 953-967 (1977).

73. Stukenberg, J. R. and Wahbeh, V. N., "Surface Aeration


Equipment: Field Performance vs Shop Performance Testing."
Proceedings Workshop Toward an Oxygen Transfer Standard,
Asilomar Conference Grounds, Pacific Grove, California,
April 11-14, 1978, pp. 163-179 (1979).

74. Benjes, H. H. and McKinney, R. E., "Specifying and Evaluating


Aeration Equipment." Journal Sanitary Engineering Division,
ASCE, 93, 55-64 (1967).

75. United States Environmental Protection Agency, Proceedings:


Workshop Toward an Oxygen Transfer Standard. Office of
Research and Development, Environmental Research Information
Center, Cincinnati, Ohio (1979).

76. Lewis, W. K. and Whitman, W. G., "Principles of Gas Absorption."


Industrial Engineering Chemistry,~ 1215-1220 (1924).

77. Higbie, R., "The Rate of Absorption of a Pure Gas into a Still
Liquid During Short Periods of Exposure." Transactions
American Institute of Chemical Engineering, .l!., 365-388
(1935).

78. Danckwerts, P. V., "Significance of Liquid-Film Coefficient in


Gas Adsorption. 11 Industrial Engineering Chemistry, ~'
1460-1467 (1951).

79. Rich, L. G., Unit Operations of Sanitary Engineering. John Wiley


and Sons, Inc., New York, New York (1961).

80. Standard Methods for the Examination of Water and Wastewater,


15th Edition, American Public Health Association,
Washington, D.C., (1980~.
143

81. Stukenberg, J. R., Wahbeh, V. N. and McKinney, R. E.,


"Experiences in Evaluating and Specifying Aeration
Equipment." Journal Water Pollution Control Federation, ~,
66-82 (1977).

82. Eckenfelder, W.W. and Ford, D. L., Water Pollution Control. The
Pemberton Press, Jenkins Publishing Company, New York, New
York (1970).

83. Wastewater Treatment Plant Design, ASCE - Manuals and Reports on


Engineering Practice - No. 36. Lancaster Press, Inc.,
Lancaster, Pennsylvania (1977).

84. Stenstrom, M. K. and Gilbert, R. G., "Effects of Alpha, Beta and


Theta Factor upon the Design, Specification and Operation of
Aeration Systems." Water Research, 15, 643-654 (1981).

85. Baillod, C.R., "Proposed Standard for Measurement of Oxygen


Transfer Rate in Clean Water." A.S.C.E. Oxygen Transfer
Standards Subcommittee, New York, New York (1982).

86. Tsao, G. T., "Simultaneous Gas-Liquid Interfacial Oxygen


Absorption and Biochemical Oxidation." Biotechnology and
Bioengineering, 10, 765-785 (1968).

87. Naimie, H. and Burns, D., "Cobalt Interference in the 'Non-Steady


State Clean Water Test' for the Evaluation of Aerator
Equipment - I. Causes and Mechanisms." Water Research, ..!..!_,
659-666 (1977).

88. Busch, A. W., "Biological Factors in Aerator Performance." Water


and Sewage Works, 117, 384-388 (1970).

89. Casey, T. J. and Karmo, O. T., "The Influence of Suspended Solids


on Oxygen Transfer in Aeration Systems." Water Research, .!!_,
805-811 (1974).

90. McWhirter, J. R., "Fundamental Aspects of Surface Aerator


Performance and Design." Proceedings 20th Purdue Industrial
Waste Conference, Purdue University, W. Lafayette, Indiana,
Extension Series 118: 75-92 (1965).

91. Otoski, R. M., Brown, L. C. and Gilbert, R. G., "Bench and Full
Scale Tests for Alpha and Beta Coefficient Variability
Determinations." Proceedings 33rd Purdue Industrial Waste
Conference, Purdue University, W. Lafayette, Indiana, Vol.
33, 835-852 (1978).
144

92. Eckenfelder, w. W. and Ford, D. L., "Engineering Aspects of


Surface Aeration Design." Proceedings 22nd Purdue
Industrial Waste Conference, Purdue University,
W. Lafayette, Indiana, Extension Series 129: 279-291
(1967).

93. Gilbert, R. G. and Chen, S. J., "Testing for Oxygen Transfer


Efficiency in a Full-Scale Deep Tank." Proceedings 31st
Purdue Industrial Waste Conference, Purdue University,
W. Lafayette, Indiana, Vol. 31, 291-311 (1976).

94. Arora, M. L., "How Much Air?" Water Engineering and Management,
129, 43-47 (1982).

95. Kayser, R., "Comparison of Aeration Efficiency Under Process


Conditions." International Association on Water Pollution
Research, Vol. 4, 477-486 (1969).

96. Novak, R. G., "Techniques and Factors Involved in Aerator


Selection and Evaluation." Journal Water Pollution Control
Federation, 40, 452-463 (1968).

97. Stenstrom, M. K., "Models for Oxygen Transfer: Their Theoretical


Basis and Implications for Industrial Wastewater
Treatment." Proceedings 33rd Purdue Industrial Waste
Conference, Purdue University, W. Lafayette, Indiana, Vol.
33, 679-686 (1978).

98. Boyle, W. C., Berthovex, P. M. and Rooney, T. C., "Pitfalls in


Parameter Estimation for Oxygen Transfer Data." Proceedins
28th Purdue Industrial Waste Conference, Purdue University,
W. Lafayette, Indiana, Vol. 28, 645-660 (1973).

99. Boyle, W. C. and Paulson, W. L., "Progress- Toward Standardized


Oxygen T.ransfer Test Procedures." Progress in Water
Technology,..!.!_, 161-170 (1979).

100. Stenstrom, M. K., Brown, L. C., and Hwang, H. J., "Oxygen


Transfer Parameter Estimation." Journal Environmental
Engineering Division, ASCE, 107, 379-397 (1981).

101. Brown, L. C. and Baillod, C.R., "Modeling and Interpreting


Oxygen Transfer Data." Journal Environmental Engineering
Division, ASCE, 108, 607-628 (1982).

102. Schultze, K. L. and Kooistra, R. D, "Oxygen Demand and Supply in


an Activated Sludge Plant." Journal Water Pollution Control
Federation, 41, 1763-1774 (1969).
145

103. Bartholomew, W. H., Karow, E. O., Sfat, M. R. and Wilhelm, R.H.,


"Mass Transfer of Oxygen in Submerged Fermentation of
Streptomyces griseus." Industrial and Engineerig Chemistry,
42, 1801-1809 (1950).

104. Benoit, R. E., Personal Communication, Blacksburg, Virgina, June


1983.

105. Krieg, N. R., Personal Communication, Blacksburg, Virginia, June


1983.

106. Gaudy, A. F. and Gaudy, E.T., Microbiology for Environmental


Scientists and Engineers, McGraw-Hill Book Company, New
York, New York (1980).

107. Brock, T. D., Biology and Microorganisms, 3rd edition,


Prentice-Hall, Inc., Englewood Cliffs, New Jersey (1979).

108. Hassan, H. M. and Fridovich, I., "Superoxide Dismutase and Its


Role for Survival in the Presence of Oxygen." in Strategy of
Microbial Life in Extreme Environments, ed. Shilo, M.,
Berlin: Dahlem Konferenzen, (1979).

109. Cooper, U. J. and Zika, R. G., "Photochemical Formation of


Hydrogen Peroxide in Surface and Ground Waters Exposed to
Sunlight, Science, 220, 711-712 (1983).

llO. Busch, A. W., "Bounds Analysis in Oxygen Transfer, In-Stream


Aeration, Solids Separation and Storm Water Problems."
American Institute of Chemical Engineering, 76, 340-343
(1979). -

lll. Bennett, C. F. and Kempe, L. L., "Oxygen Transfer Mechanisms in


the Gluconic Acid Fermentation by Pseudomonas ovaus."
Biotechnology and Bioengineering,~, 347-360 (1964).

ll2. Bennett, G. F. and Kempe, L. L., "Oxygen Transfer in Biological


Systems." Proceedings 20th Purdue Industrial Waste
Conference, Purdue University, W. Lafayette, Indiana, Vol.
20, 435-449 (1965).

ll3. Albertson, O. E. and DiGregorio, D., "Biologically Mediated


Inconsistencies in Aeration Equipment Performance." Journal
Water Pollution Control Federation,£_, 976-988 (1975).

ll4. Eckenfelder, W. W., "Oxygen Transfer: A Historical


Perspective - The Need for a Standard." Proceedings
Workshop Toward an Oxygen Transfer Standard, Asilomar
Conference Grounds, Pacific Grove, California, April 11-14,
pp. 1-2 (1979).
146

115. United States Environmental Protection Agency, Manual of Methods


for Chemical Analysis of Water and Wastes, Office of
Technology Transfer, Washington, D.C. (1974).

116. Association of Environmental Engineering Professors,


Environmental Engineering Unit Operations and Unit Processes
Laboratory Manual, O'Connor, J. T. (editor) (1972).

117. Sherrard, J. H., Schroeder, E. D. and Lawrence, A. W.,


"Mathematical and Operational Relationships for the
Completely Mixed Activated Sludge Process," Water and Sewag~
Works, 121, 84-91 (1974).

118. Hoover, S. R. and Porges, N. "Assimilation of Dairy Wastes by


Activated Sludge." Sewage and Industrial Wastes, 24,
306-312 (1952).

119. Sherrard, J. R. and Benefield, L. D., "Use of Elemental


Distribution Diagrams in Biological Wastewater Treatment."
Journal Water Pollution Control Federation, 48, 562-569
(1976). -

120. Andrews, J. F. and Graef, S. P., "Dynamic Modeling and Simulation


of the Anaerobic Digestion Process." in Anaerobic Biological
Treatment Processes, Advances in Chemistry Series 105, R. F.
Gould, ed., American Chemical Society, Washington, D. C.
(1971).

121. Haldane, J. B. S., Enzymes. The M.I.T. Press, Massachusetts


Institute of Technology, Cambridge, Massachusetts (1965).

122. Metcalf and Eddy, Inc., Wastewater Engineering: Collection,


Treatment, Disposal. McGraw-Hill, Inc., New York, New York
(1972).

123. Mines, R. o. and Sherrard, J. H., "Activated Sludge Biokinetic


Constant Evaluation," in press, Journal Water Pollution
Control Federation.

124. Barnhart, E. L., Personal Communication, Ithaca, New York, June


1983.
APPENDIXA

LISTINGS OF THE FORTRAN


ANDSAS

COMPUTER
PROGRAMS
USED IN THE MODELING
STUDIES

147
C STOICHIOMETRIC PROGRAM
C
C ORIGINAL PROGRAM BY ADIL GODREJ, MODIFIED BY HICIIARD O. MINES
C
REAL CNR( 10), THETAC( 10), SOCOD( 10), SlCOD( 10), ECODR( 10), SlN I TS( 10),
1 SONITS( 10), E FFNIT ( 10) , ENH4R( 10), E FN02N ( 10), E FN03 N( 10), 03 ( 10),
2 YOCOD(10),YONITS(10),YONITB(10),XCOD(10),XNITS(10),XNITB(10),
3 CODSP( 10),NITSSP( 10),NITBSP( 10),NITSP( 10). TOTSP( 10),XNIT( 10),
11 C5N( 10), NHl1N( 10). N02N( 10), N03N( 10), 02REQD( 10), 02RQNS( 10),
5 02RQNB(10),02UPTK(10),02UPN1(10),SONH4(10),G2(10),C6C(10),
6 SIN ITS(10), EFN02( 10), EFN03( 10), SONI TB( 10), SlN I TB( 10), El ( 10),
7 TOTALN( 10), TOTALC( 10) ,ALKCliG( 10), BIi( 10), C5C( 10), C02C( 10),
8 A 1 ( 10), B 1 ( 10) , D1 ( 10), C 1 ( 10), A2 ( 10) , B2 ( 10), YON IT ( 10), B5 ( 10),
9 Gl ( 10), E2( 10), D2( 10), C2( 10), B3( 10), 02RQN I ( 10) ,ARATIO( 10 ),
A F2( 10), E3( 10), F3( 10), D3( 10 ), C3( 10), B6( 10 ),XMLSS( 10), Fl ( 10),
B NRATI0(10),02COEF( 10),V,YMCOD,KDCOD,KCOD,KSCOD,G2G3,
C KNITS,KSNITS,YMNITB,YMNITS,KDNITS,KNITB,KDNITB,KSNITB,AMMON,
D LMI 10,CONC,MW,MWORGS,THETA,G1G2
READ (5,10) N,CONC,MW,Q,V,MWORGS,AMMON,LMITO
10 FORMAT( 12, F8. l, F5.0, F10.1, F5.3,6X, F5.0, F8.1, Fl0.4)
READ (5,20) YMCOD,KDCOD,KCOD,KSCOD,YMNITS,KDNITS,KNITS,KSNITS,
lYMNITB,KDNITB,KNITB,KSNITB
20 FORMAT (8F10.3/4f10.3)
READ (5,30) (TIIETAC(l),1=1,10)
30 FOHMAT ( 10F5. 1)
TIIETA=V/Q
WHITE (6,401
1,0 FORMAT (lX, SUMMARY OF PARAMETERS USED')
WRITE (6,50) N,CONC,MWORGS,Q, THETA,AMMON
50 FORMAT( '-NUMBER OF OC VALUES USED=' 12/lH+ T12, 1 - 1 /'0CONCENTRATIO
1N OF SUBSTRATE (GLUCOSE)= ',F8.1 1X, 1MG/L'/ 10MOLECULAR WEIGHT OF M
21CROORGANISMS (C5H702N)= 1 ,F5.0/ 10FLOWRATE= 1 ,Fl0.1, lX, 'LITERS/DAY
3'/'0HYDRAULIC RETENTION TIME=' 12X,F5.3, lX, 'DAY'/'OINFLUENT NH4-N
4CONCENTRATION =',F8.1,1X, 1 MG/L )
WRITE (6,60) YMCOD,KDCOD,KCOD,KSCOD,YMNITS,KDNITS,KNITS,KSNITS,
lYMNllB,KDNITB,KNITB,KSNITB
60 rDRMAT( '-BIOKINETIC CONSTANTS USED:' /'OltETEROTROPIIIC REACTION 1coo
1 REMOVAL):'/' YMAX=',F10.3,1X,'MG VSS/MG COD'/' KD==' F10.3,1X, /DA
2Y 1 / 1 K=',Fl0.3,lX, '/DAY'/' KS=',Fl0.3,lX, 'MG/L COD'/ 10NITROSOMONAS
3 HlACTION (NHl1 OXIDATION):'/' YMAX= 1 ,f10.3,1X,'MG VSS/MG Nlll1-N'/, 1

4 KDc.c' F10.3,1X,'/DAY 1/ 1 K=',Fl0.3,lX, 1 /DAY'/' KS=',F10.3,1X, 1 MG/L

5Nlll1-N 1/ 1 0NITROBACTER REACTION (N02 OXIDATION):'/' YMAX= 1 ,F10.3 lX


6' VSS/MG N02-N' /' KD:=' , F 10. 3, 1X, '/DAY'/' K=', F 10. 3, 1X, '/DAY' / 1 KS= I
7,F10.3,1X,'MG/l N02-N 1 )
WRITE (6 70) (THETAC(l),1=1,10)
70 FORMAT( 1-MEANCELL RESIDENCETIMES USED (DAYS): 1 /'+-'/1H0,10F8.1)
C THE EQUATIONSARE SET-UP AS FOLLOWS:
C
C HETEROTROPHIC
REACTION(COD REMOVAL):

C
C Al C6ll1206 + Bl NH4+ + Dl 02 ---> Cl C511702N+ A2 C6H1206 + 82 N114+
C + El CO2+ Fl H20 + Gl H+
C
C NITROSOMONAS
REACTION(NH4 OXIDATION):
C
C E2 CO2 + 82 N114++ D2 02 ---> C2 C5H702N + 83 NHl1++ 84 N02- + G2 H+
C + F2 H20
C
C NITROBACTERREACTION(N02 OXIDATION):
C
C E3 CO2 + BIi N02- + F3 1120 + D3 02 + G3 II+ ---> C3 C5H702N + 85 N02-
C + 86 N03-
C
COVERALL BALANCED
STOICHIOMETRICEQUATION:
C
C Al C6111206 + Bl NH4+ + (Dl+D2+D3) 02 ---> (Cl+C2+C3) C5H702N
C + A2 C6H1206 ¥ 83 NH4+ + 85 N02-
C + 86 N03- + (Gl+G2-G3) H+
C + (Fl+F2-F3) 1120 + (E1-E2-E3)C02
DO 300 1=1,N
SOCOD(I )'-"CONC
Al( i)=CONC/MW
81( I )=AMMON/14.0
SONll4( I ) =AMMON
S1COD(I )=(KSCOD*(l+KOCOD*THETAC( I )))/(THETAC( I)*
l(YMCOD*KCOD-KDCOD)-1.0)
CNR(l)=SOCOD(1)/AMMON
A2( I )=Al( I )*SlCOD( I )/SOCOD(I)
YOCOD(I )=YMCOD/(1.0+KDCOD*THETAC( I))
Cl( l)=YOCOD(I )*(SOCOD(l)-SlCOD( 1))/MWORGS
El( I )=6.0*Al( I )-5.0·H-Cl( I )-6.0*A2( I)
82( I )=Bl ( I )-Cl( I)
Gl( I )=Bl( I )-82( I)
Fl( I )=(12*Al( I )+l1*Bl( I )-(7*Cl( I )+12•11-A2( I )+11*82( I )+Gl( I )))/2.0
01 ( I )= ( 2*C1 ( I )+6*A2 ( I )+2*E1 ( I )+F 1( I )-6*A 1( I ) )/2. 0
COOSP( I )=YOCOD( I )"Q*(SOCOD( I )-SlCOD( I) )/453000.0
SONI TS( I )=1lt.O"B2( I)
SlN I1S( I)=( KSN I1S*( 1. O+KDN I TS*THETAC( I)))/( THETAC( I)*( YMN I TS*KN I TS
1-KDNITS)-1.0)
If( (SONI TS( I )-SlNITS( I) ).GT.0.0.AND.SlNITS( I ).GT.O.O)GOTO 80
75 SONITS(l)=O.O
SlNllS( I )=0.0
83( I )-=02( I)
YONI TS( I )=0.0
C2( I )=O. 0
E2( I )=0.0
BIi( I )=0.0
F2( I )=O.O
D2(I)=0.0
G2( I )=0.0
NITSSP( I )=0.0
GO TO 90
80 B3( I )=B2( I )*SINITS( I )/SONI TS( I)
YON ITS~ I )=YMNITS/(1.0+KDNITS*lHETAC( I))
I-'
V,
0
NITSSP( I )=YON ITS( I )*Q'"·(SONITS( I )-SlNITS( I) )/453000.0
C2 ( I ) =YON I TS ( I ) * ( SONI TS( I )-S 1 NI TS( I ) ) /MWORGS
E2( I )=5. O*C2( I)
BIi( I )=82( I )-C2( I )-83( I)
G2( I )=82( I )-B3( I )+B4( I)
F2( I )=(11*82( I )-(7*C2( I )+4*83( I )+G2( I)) )/2.0
D2( I }=(2.0*C2( I )+2.0*B4( I )+F2( I )-2.0"'E2( I) )/2.0
SONI TB( I )=14.0"B4( I)
S 1N118( I )-= ( KSN I TB*( 1. O+KDN I TB*THET AC( I ) ) ) / ( THETAC( I ) *( YMN I TB*KN I TB
1-KDNITB)-1.0)
IF( (SONI TB( I )-SlNITB( I) ).GT.0.0.AND.S1NITB( I ).GT.O.O)GOTO 100
90 SONIT8( l)=O.O
SlNITB( I )=0.0
G3( I )=0.0
B6( I )=O. 0
8 5 ( I ) = 8/1 ( I )
YONI TB( I )=0.0
C3(I)=0.0
[)3(I)=0.0
F3(I)=0.0
NI lBSP( I )=0.0
[3(I)=0.0
GOTO 110
100 85( I )=-S1Nllll( I )/1ll,0
YONI TB( I )=YMNI TB/( 1.0+KDNITB*lllETAC( I))
NI TBSP( I )=YON I TB( I )*Q*(SONITB( I )-SlNITB( I) )/453000,0
C3( I )=YON I TB( I)*( SONI TB( I )-S1 NI TB( I ) )/MWORGS
86( I )=BIi( I )-C3( I )-B5( I)
G3( I )=-BIi( I )-85( I )-86( I)
F3( I )=(7*C3( I )-G3( I ))/2.0
E3( I )=5*C3( I)
D3( I )=(2*C3( I )+3*B6( I )+2*85( I )-(2*E3( I )+2*Bll( I )+F3( I)) )/2.0
EFN03( I )=86( I )*1II,0
11 0 f.CODR( I ) = ( ( SOCOO( I ) - S1COO( I ) ) * 100 . 0 ) / SOCOD( I )
XCOO( l)=(YMCOD*(SOCOO( I)-SlCOD( l))*THETAC( I))/((1.0+KOCOO*THETAC( I
l))*TlffTA)
XN ITS( I)=( YMN I TS*( SONITS( I )-S1 NI TS( I) )*TllET AC( I ) )/ ( ( 1. O+KON I TS*THE
lTAC( I) )*TIIETA)
XN 11 B( I )=-( YMNI TB*( SONI TB( I )-Sl NI TB( I ) ) *THETAC( I ) ) / ( ( 1. Q+KON I TB*THE
lTAC( I ) ) *HIETA)
D102=D1 ( I )+D2( I)
G1G2=G1 ( I )+G2( I) I-'
V,
G2G3=G2( I )-G3( I)
I-'
G1G2G3~G1( l)+G2G3
IF(SlNITS( I ).LE.0.0.AND.SlNITB( I ).LE.O.O)ALKCIIG( I )=-50.0*Gl( I)
IF( SlN I TS( I). GT. 0. O.AND. SlN I TB( I). LE. U. 0 )ALKCIIG( I )=-50. Q·H-G1G2
IF(S1NITS( I ).GT.0.0,AND.SlNITB( I ).GT.O.O)ALKCHG( I )=-50.0*G1G2G3
D1D2D3=D3( I )+0102
IF(I.MITO.GE.01D203)GOTO 160
IF(LMITO.GT.D102.AND.LMITO.LT.D1D2D3)GOTO 140
02COEF( I )=LM I TO-D1 ( I)
WRI TE( 6, 120) LM I 10, 02COEF ( I ) , D2 ( I )
120 FORMAT( 1 lAERATION CAPACITY RESTRICTS TIIE AMOUNT OF OXYGEN THANSF
!ERRED TO 1 ,F10.4,1X, 1 M-MOLES OF OXYGEN PER DAY 1 / 1 THE AMMONIUM REA
2CTION IS ONLY SUPPLIED •,~10.4,lX,'M-MOLES OF OXYGEN PER DAY'/' CO

3MPLEfE AMMONIUM OXIDATION REQUIRES ',F10.4,1X,'M-MOLES OF OXYGEN P


l1ER DAY')
ARATIO( I)=-02COEF( I )/02( I)
C2( I )=C2( I )*ARATIO( I)
[2( I )=[2( I )*AHATIO( I)
BIi( I )=BIi( I )·11-ARAl10( I)
H3( I )=B2( I )-(C2( I )+BIi( I))
G2( I )=-G2( I )*ARATIO( I)
F2( I )=F2( I )·K·ARATIO( I)
D2( I )=02COEF( I)
YON I TB( I )=0.0
G1G2=G1 ( I )+G2( I)
C3( I )=0.0
85( I )=BII( I)
B6( I )=0.0
[3(1)=0.0
F3(I)=0.0
03(1)=0.0
G3(1)=0.0
NI TBSP( I )=0.0
SONI TB( I )=0.0
XNITB( I )=0.0
ALKCHG( I )=-50.0*G1G2
SlNITS( I )=O.O
SlNITB( I )=0.0
SlNITS( I )=14.0*BII( I)
YON ITS( I )=C2( I )*MWORGS/( SONI TS( I )-SlNITS( I))
NITSSP( I )=YON ITS( I )*Q*(SONITS( I )-SlNITS( I) )/453000.0
XN I TS( I )=YON I TS( I)*( SONI TS( I )-S1 NI TS( I ) )*HIETAC( I )/HIET A
GOTO 170
JII() 02COEF( I )=LMITO-D102
1115 WRITE(6, 150)LMIT0,02COEF( I ),D3( I)
150 FORMAT( 1 1AERATION CAPACITY RESTRICTS THE AMOUNT OF OXYGEN TRANSF
lERRED TO 1 ,F10.4,1X, 1 M-MOLES OF OXYGEN PER DAY 1 / 1 THE NITRITE REAC
2TION IS ONLY SUPPLIED 1 ,F10.4,1X, 1 M-MOLES OF OXYGEN PER DAY 1 / 1 COM

JPLETE NITRITE OXIDATION REQUIRES' ,F10.4,1X, 1 M-MOLES OF OXYGEN PER

ltOAY 1 )
NRAT 10( I )=02COEF( I )/D3( I)
C3( I )=C3( I )*NHATIO( I)
E3( I )=E3( I )*NRATIO( I)
86( I )=B6( I )*NHAT I 0( I)
85( I )=B4( I)-( C3( I )+B6( I))
G3( I )=G3( I )*NRATIO( I)
F3( I )ccf3( I )·K·NHATIO( I)
1>3( I )=02COEF( I)
G2G3=G2( I )-G3( I)
G1G2G3=G1( I )+G2G3
ALKCHG( I )=-50.0*G1G2G3
SlNITB( I )=ll1,0KB5( I) ·
YON I TB( I )=C3( I )*MWOHGS/( SON I TB( I )-SlN I TB( I))
NI TBSP( I )=YON I TB( I )*Q""( SON I TB( I )-S 1N I TB( I) )/1153000. 0
XN I TB( I )-=YONI TB( I )*(SONI TB( I )-Sl NI TB( I ) )*HIET AC( I )/TIIET A
GOTO 170
160 WRITE(6 I 165)LMITO
165 FORMAT( lAERATION CAPACITY RESTRICTS THE AMOUNT OF OXYGEN TRANSFER

lLO TO ',Fl0.I1,lX,'M-MOLES Of OXYGEN PER DAY'/' OXYGEN DEMAND DOES


2NOT EXCEED TRANSFER CAPACITY 1 / 1 THEREFORE THE SYSTEM IS ONLY CARB
30N LIMITED AND NOT OXYGEN LIMITED')
170 EHNI f( I )=14.0*·BJ( I)
ENIIIIR( I)=( ( SONII4( I )-EFFN IT( I) )*100. 0 )/SONH4( I)
EFN02( I )=1lt.OMB5( I)
EFN03( I )=86( I )*14.0
XMLSS( I )=XCOD( I )+XN I TS( I )+XN I TB( I)
XNIT( I )=XNITS( I )+XNITB( I)
TOTALN( I )=Cl( I )+C2( I )+C3( I )+B3( I )+B5( I )+B6( I)
C5N( I)=( (Cl( I )+C2( I )+C3( I) )*100.0)/TOTALN( I)
Nlll1N( I )=B3( I )*100.0/TOTALN( I)
N02N( I )=135( I )*100.0/TOTALN( I)
N03N( I )=136( I )*100.0/TOTALN( I) I-'
V,
IJ1D2D3=01 ( I )+IJ2( I )+D3( I) w
NITSP( I )=NITSSP( I )+NI TBSP( I)
TOT SP( I ):ccCODSP(I )+NI TSP( I )
C1C2C3=C1 ( I )+C2( I )+C3( I)
SC2C3=C2 ( I ) +C3 ( I )
Sl2E3=E2( I )+E3( I)
YONIT( I )=YON ITS( I )+YON I TB( I)
SD2D3=02( I )+03( I)
f2MF3=F2( I )-f3( I)
ElM=El ( I )-SE2E3
F 1M=F1 ( I )+F2MF3
TOTALC( I )=5*( C1C2C3 )+6·*A2( I )+ElM
C'.,C( I )c:c(5.0*C1C2C3*100.0)/TOTALC( I)
C02C( I )=(El( I )-SE2E3)*100.0/TOTALC( I)
C6C( I )-=6*A2( I )*100.0/TOTALC( I)
02REQD( I )=32.0*010203*Q/453000.0
02RQNS( I )=32.0*02( I )*Q/453000.0
02RQNB( I )=32.0*D3( I )*Q/lI53000.0
02HQN I ( I )=32. OitDl ( I )*Q/453000. 0
02Ul'TK( I )=02REClD( I )*lI53000.0/(XMLSS( I )*V)
0?.IJl'N I ( I )=02HQN I (I) 111153000. 0/( XCOD( I) *V)
C DATA OUTPUT TO PHINT FILE BY TIIETAC VALUES
WRITE(6,200)THEfAC( I)
200 FORMAT ( 1 0FOR QC= '/'+ 1 ,F~.1, 1 DAYS:')
WRI TE(6,210 )CNR(I), SOCOD(I), SlCOD( I ) , ECODR(I ) , YOCOD(I), XCOD(I), A
11 ( I ) , B1 ( I ) , D1( I ) , C1 ( I ) , A2( I ) , B211 ) , E1( I ) , F1( I ) I G1 ( I )
210 FORMAT( 1 -COD:TKNRATIO=',F7.2/ INFLUENTCOD= ,F8.2 lX,'MG/L'/ 1 E
lEFLUENT COD=1 ,F8.2,1X, 'MG/L1 / 1 COD REMOVAL EFFIENCY=I1 F6.2,1X, 1 %1 /
2 1 Y(OBS) FOR COD REMOVAL=',F5.3,1X,'MG VSS/MG COD'/ AERATIONBA
3SIN VSS FOR COD REMOVAL=',F10.2,1X 'MG/L1 / 1 0THE BALANCED STOICHI
l10METRICEQUATIONFOR COD REMOVALIS: 111110,n.1,, 1 C6H1206 +',n.11, 1
5 Nlllt+ +',F7.lt, 1 02 --->1 ,F7.4
1 1 C5H702N + ,F7.4,' C6H1206 + ,F7.4,
6' NH4+ + , F7. 4, ' CO2 +' , F7. 4, H20 +' , F7. 4, ' H+ 1 )
IF(SONITS( I ),EQ.0.0.AND.SONITB( I ),EQ.O.O)GOTO 260
215 WRITE(6,220)SONITS( 1),SlNITS( I ),EFFNIT( I ),YONITS( I ),XNITS( I),
1NI TSSP( I ), E2( I ) , B2( I ) , D2( I ) , C2( I ) , 83 ( I ), B4( I ) , G2( I 1,F2 ( I )
220 FORMAT('-SO FOR AMMONIUM OXIDATION=',F8.2,1X, 'MG/L / 1 S1 FOR AMMON
llUM OXIDATION= 1 ,F8.2,1X,'MG/L 1 / 1 NH4-N IN EFFLUENT=',F8.2,1X 'MG/L
2 1 / 1 Y(OBS) FOR AMMONIUM OXIDATION=',F5.3 lX,'MG VSS/MG NH4-N1/' AE
3RATION BASIN vss FOR AMMONIUM OXIDATION= 1,F10.2,1X, 1 MG/L / SLUDGE
1 1

4 PRODUCTION 1 ,F10.4,1X,'LB/DAY 1 / 1 0THE BALANCED STOICHIOMETRICEQUAT

510N FOH AMMONIUM OXIDATIONIS:'/1HO,F7.4,' CO2 +',F7.4,' N114++' F


67.4, 1 02 ---> 1 ,F7.4,' C5H702N + 1 ,F7.4, 1 NH4+ +',F7.4, 1 N02- +',F~.
71,,' II++' ,F7,lt,' 1120 1 )
IF(SONITS( l),GT.0.0.AND.SONITB( I ).EQ.O.O)GOTO 270
WRITE(6,230)SONI TB( I), SlNITB( I), SlNI TB( L), YONITB( I) ,XNITB( I),
1NI TBSP( I ) , E3( I ) , B4( I ) , F3( I ) , D3( I ) , G3( I ), C3( I , , B5( I ) I B6( I )
230 FORMAT( '-so FOR NITRITE OXIDATION=',F8.2,1X, MG/L'/ Sl FOR NITRIT
1E OXIDATION= 1 ,F8.2,1X, 1 MG/L1 / 1 N02-N IN EFFLUENT=',F8.2,1X 'MG/L'/
2 1 Y(OBS) FOR NITRITE OXIDATION=',F10.3,1X,'MG VSS/MG N02-N1/ 1 AERA
3TION BASIN VSS FOR NITRITE OXIDATION=',F10.2,1X, 'MG/L'/ 1 SLUDGEPR
40DUCTION=',F10.4,1X,'LB/DAY 1 / 1 0THE BALANCED STOICHIOMETRICEQUATIO
5N FOR NITRITE OXIDATIONIS:'/1HO,F7.4 1 ' CO2 +',F7.4,' N02- +',F7.4
6, 1 1120 +',F7.4,' 02 +',F7.4,' H+ ---> ,F7.4,' C5H702N +',F7.4, 1 NO
72- +', F7. 4, ' N03-' )
235 WRITE(6,21tO)SONITS( I), SlNITS( I), EFFNIT( I), YONIT( I ),XN IT( I),
lNITSP( I ,,SE2E3,B2( I ),SD2D3,SC2C3,B3( I ),B5( I ),B6( I ,,G2G3, F2Mf3
21,0 fORMATjOSO FOR NITRIFICATION=',F8.2,1X, 'MG/L',/, Sl FOR NITRlflC
lATION= ,F8.2,1X,'MG/L 1 / 1 NH4-N IN EFFLUENT=',F8.2 1X,'MG/L 1 / 1 Y(OB
2S) FOR NITRIFICATION==',F5.3,1X, MGVSS/MG Nlll1-N ;i' AERATIONBASIN
1 1 1
3VSS FOR NITRIFICATION=1 ,Fl0.2,1X,'MG/L 1 / 1 SLUDGEPROD. NIT. = 1 ,FlO
4.4,lX,'LB/DAY'/'OTHE BALANCED STOICIIIOMETRICEQUATIONFOR NITRIFIC
5ATION IS: 1 /1110,F7.1t,' CO2 + 1 ,F7.1i, 1 NHlt+ + 1 ,F7.4, 1 02 ---> 1 ,F7,Lt,'
6 C51l702N +',f7.lt,' N114++',F7.4,' N02- +',F7.IJ, 1 N03- +' F7.4, 1 H+
7 +' , F7. Lt, ' 1120' ) '
WRITE(6,250)Al( I ).Bl( I ),D1D2D3,C1C2G3,A2( I ),B3( I ).EIM,85( I ),86( I),
1G1G2G3,F1M
250 FORMAT('OTHEBALANCED BIOCHEMICALSTOICHIOMETRICEQUATIONFOR THE
lTOTAL REACTIONIS: 1 /lll0 1 F7,11,' C6111206 + 1 1 n.11, 1 Nlll1+ + 1 ,F7,l1,' 02
2 ---> 1 ,F7.4,
3C02 +' ,F7.4,'
1 C5H702N + ,F7.4r'
N02- '/'O',T81,
C6H1206 + ,F7.4,' NH4+ +',F7.4,'
+',F7.4, 1 N03- +',F7.4, 1 H+ +',F7.4
lj
I I 1120I )
GOTO276
260 WHI TE.( 6 265 )
265 FORMAT( 1oTHERE IS NO NITRIFICATION SO THE BALANCED STOICIIIOMETRIC
lEQUATIONFOR'/' THE TOTALREACTIONIS TIIE SAMEAS ntAT FOR THE CAR
280N REMOVAL EQUATION.')
GOTO2115
270 WRITE(6 275)
275 FORMAT( 10PARTIAL AMMONIUM OXIDATIONIS OCCURRINGBUT NITRITE OXIDA
lTION IS INHIBITED'/' THE TOTALREAC[ION IS THE SAMEAS THAT FOR
2COD REMOVAL ANDAMMONIUM OXIDATION.')
GOTO235
276 WRITE(6 277)C5C( I ),C02C( I ),C6C( I ),CODSP( I)
277 FORMAT( 10THE PERCENTDISTRIBUTIONOF CARBONIS:'/' SYNTHESIZEDCAR
lBON (C5H702N)= 1 ,F6.3,1X, '%'/' CO2 CARBON=',F6.3,1X '%'/' GLUCOSEC I-'
2ARBON(C6111206)=',F6.3,1X,'%'/' SLUDGEPRODUCTION= 1 ,no.4,1X,'LB/D V,
V,
3AY1 )
278 WRI TE( 6,279 )C5N( I), NH4N(I), N02N( I), N03N( I), TOTSP(I)
279 FORMAT( 'OTHE PERCENTDISTRIBUTIONOF NITROGENIS:'/' SNYTHESIZEDN
11 TROGEN='.F6. 3, 1x, '%' /' Nll4-NI TROGEN=',F6. 3 1ix,'%'/' N02-N I rnoGEN=
2',F6.3,1X,'%'/' N03-NITROGEN= 1 ,F6.3,1X,'%'/ TOTALSLUDGEPRODUCT!
30N ( INCLUDESCOD REM.)=',F10.4,1X, 'LB/DAY')
280 WRIrE(6,285)02REQD( I ),02RQNS( I ),02RQNB( I ),02RQNI ( I ),02UPTK( I ),02UP
1NI ( I ),ALKCHG(I)
285 FORMAT( 'OTOTALOXYGENREQUIREMENTS ( INCLUDESNITRIFICATION)=',F10.

111,lX,'LB OF 02/DAY'/' OXYGENREQUIREMENTS FOR NH4+10N OXIDATION=',


2r10.11,1X, 1 LB OF 02/DAY1 / 1 OXYGENREQUIREMENTS FOR NITRITE OXIDATIO
3N=',FI0,4 lX, 'LB Of 02/DAY'/' OXYGENREQUIREMENTS(NITRIFICATION IN
4HIBITED)= 1,F10.4,1X,'LB OF 02/DAY'/' SPECIFIC OXYGENUPTAKE(INCLUD
5ES NITRIFICATION)=1 ,f10.4,1X, '/DAY'/' SPECIFIC OXYGENUPTAKE(NITRI
6f ICATION I NIII BI TED)=', flO, 4, 1X, '/DAY' / 1 ALKALINITYCHANGE= 1 , F9. 3, 1

7X, 1 MG/L AS CAC031 )


C WRITE NEEDEDOUPUTTO FILE FOR FURTHERPROCESSING
WRITE(7,290)THETAC( l),SOCOD( I ),S1COD( I ),ECODR(I ),SONH4( I),
l[FFNIT( I ),ENH4R( I ),EFN02( I ),EFN03( I ),YOCOD(I ),YONITS( I ),YONITB( I),
2XCOD(I ),XNITS( I ).XNITB( I ),XMLSS( I ),C5C( I ),C02C( I ),C6C( I ),C5N( I),
3NIIIIN( I), N02N( I ). N03N( I), 02REQD( I ) , 02RQNS( I). 02RQNB( I ) ,
l1ALKCHG( I). CODSP( I), NI TSSP( I ) , NI TBSP( I ) , TOT SP( I)
290 FORMAT(6(5F15.5,/),F15.5)
300 CONIINIJE
C OUPUT DATA IN TABULAR FORM
WHITE ( 6, 3 10) ( CNH( I ) , I"' 1 , N), ( THET AC ( I ) , I= 1 , N), ( SOCOD( I ) , I= 1 , N), ( S 1 C
100 ( I ) , I= 1 , N) , ( ECODR( I ). I= 1 , N ). ( SONIIII( I ). I= 1 , N) , ( E FF N I T ( I ). I= 1 , N ) ,
2 ( lNIIII R( I ) , I= 1, N), ( EFN02 ( I ) , I==1 , N)
WHITE ( 6, 320) ( HN03 ( I ) , I= 1 , N), ( YOCOD( I ) , I= 1, N), ( YON ITS( I ) , I= 1 , N), (
1YON IT B ( I ) , I = 1 , N ) , ( XCOD( I ) , I= 1 , N), ( XN ITS ( I ) , I= 1 , N) , ( XN I TB ( I ) , I= 1 , N)
2, ( XMLSS ( I ) , I= 1 , N) , ( C5C ( I ) , I= 1 , N ) .
WRI TE( 6,330) ( C02C( I). I= 1, N), ( C6C( I), I= 1, N), ( C5N( I), I= 1, N), ( NH4N( I )
1, l:c1,N),(N02N( I), 1=1,N),(N03N( I), 1=1,N),(02REQD( I), 1=1,N),(02RQNS(
2 I ) , I= 1 , N), ( 02RQNB ( I ). I= 1 , N)
WRITE(6,3110)(02UPTK( I), 1=1,N),(02UPNI( I), 1=1,N),(ALKCIIG( I), lc::1,N),
l(CODSP( I), 1=1,N).(NITSSP( I), 1=1,N),(NITBSP( I), 1=1,N),(TOTSP( I ).1=1
2,N)
310 IORMAT( 'lCOD: TKN RAT 10 1 , T34, 10F10. 2/ 10-C (DAYS)', T33, 10F10. 1 /'
1+0' ,/'OINFLUENT COD (MG/L)' ,T34, 10F10.2/'0EFFLUENT COD (MG/L)', T34
2,10F10.2/'0COD
3) 1 ,T34,10F10.2/'0EFFLUENT
REMOVAL EFF.(%)' ,T34,10F10.2/'0INFLUENT
NH4-N (MG/L)',T311,10F10.2/
NH4-N (MG/L
10NH4-N REMOVA
....
V1
l1L EFF. \%l',T31i,10F10.2/ 10EFFLUENT N02-N (MG/L)',T34,10F10.2) 0-,

320 FORMAT( OEFFLUENT N03-N (MG/L)',T34,10F10.2/ 10Y(OBS) FOR COD REMOV


1AL 1 ,T34,10F10.3/ 10Y(OBS) FOR NH4 OXIDATION',T34,10F10.3/'0Y(OBS) F
20R N02 OXIDATION',T34,10F10.3/ 10VSS FOR COD REMOVAL',T34,10F10.2/
3'0VSS FOR NH4 OXIDATION 1 ,T34,10F10.2/'0VSS FOR N02 OXIDATION',
4134,10F10.2/'0TOTAL VSS(AERATION BASIN) 1 ,T34,10F10.2/ 1 0SYNTHESIZED

5CARBON \%l',T34,10F10.21
330 FOHMAT( OC02 CARBON(%) ,T34,10F10.3/'0GLUCOSE CAHBON (%)',T34,10F
110.3/'0SYNTHESIZED NITROGEN (%1',T34,10F10.3/'0NH11-NITROGEN (%1'•
21311,10F10.3/ 10N02-NITROGEN (%) ,T311,10F10.3/'0N03-NITROGEN (%),
3T34,10F10.3/'0TOTAL 02 REQUIRED (LB/DAY) 1 ,T34,10F10.4/'002 REQD. N
1,111,OXID.(LB/DAY) 1 ,T311,lOF10.l1/'002 REQD. N02 OXID.(LB/DAY)',134,10
5F10. It)
JL1ll FORMAT('OSPEC. 02 UP.(WITII NIT) (/DAY) 1 ,T34,10F10.4/ 1 0SPEC. 02 UP.
l(W/0 NIL) (/OAY)',T34,10F10.4/'0ALK CHANGE (MG/LAS CAC03)',T34,1
20F10.3/ 10SLUDGE PROD.(COO REM) (LB/DAY)',T34,10F10.4/ 10SLUDGE PROD
3. (N114 REM) (LB/DAY) 1 ,T31t,10F10.11/'0SLUDGE PROD. (N02 REM) (LB/DAY
4)',T34,10F10.4/'0TOTAL SLUDGE PROD.(LB/DAY) 1 ,T34,10F10.4)
STOP
END
GOPTIONS DEVICE=VPISASGV COLORS=(BL BL Bl. BL) HSIZE=7 VSIZE=9.5;
DATA FULL;
INPUT HIETAC 1-15 SOCOO 16-30 SlCOD 31-115 ECODR116-611 SONll1161-75
#2 EHNIT 1-15 EN1111R 16-30 EFN02 31-45 EFN03 46-60 YOCOO61-75
#3 YONITS 1-15 YONITB 16-30 XCOD 31-45 XNITS 46-60 XNITB 61-75
/Ill XMLSS l.:--15 C5C 16-30 C02C 31-45 C6C 46-60 C5N 61-75
//5 NIIIIN 1-15 N02N 16-30 N03N 31-115 02REQD 116-60 02RQNS 61-75
#6 02RQNB 1-15 ALKCIIG16-30 CODSP 31-45 NITSSP 46-60 NITBSP 61-75
#7 TOTSI' 1-15;
LABEL THEIAC=' MEANCELL RESIDENCE TIME (DAYS)';
CARDS;
2.00000 322.01587 38.39171 88. 07765 493. 10010
385.9211% 21. 73503 93. 4137611. 0.0 0.34381
0.033116 0.0 1165. 4119116 61. 89680 0.0
527. 311619 l18. 58092 39.49675 11.92230 2. 775811
78.26501 18.95917 0.0 0.01581 o.01067
o.o -716.651105 o. 003411 o. 000116 0.0
().00390
2.20000 322.01587 34.36160 89.32913 493. 10010
387. 04102 21.50862 92.35757 0.0 0.34032
I-'
0.03266 0.0 5111.00928 66.64722 0.0 V,
580. 656119 118.62912 40.70007 10.67078 2.77860 -..J

78. 1191112 18.72998 0.0 0. 01581 0.01055


0.0 -708.63013 0. 003116 0.00045 0.0
0.00391
3.00000 322.01587 24. 553511 92. 371198 1193. 10010
390. 1111811 20.76357 88.87123 0.0 0.32707
0.03004 0.0 696.59570 84.36011 0.0
780.95581 47. 96286 1111.41216 7. 621195 2.74053
79.23656 18.02298 0.0 0.01581 0.01017
0.0 -683.05762 0.00344 0. 000112 0.0
ll.00385
11;30000 322.01587 17. 300115 94.62741 493. 10010
395.35815 19.82191 84.81149 0.0 0.30761
0.02681 0.0 961. 92969 Hl9. 123119 0.0
1071. 05298 45.89258 48.73485 5.372511 2.62223
80.17813 17. 19966 0.0 0.01581 0.00974
0.0 -651.97632 0.00331 0.00038 0.0
0.00369
5.00000 322.01587 15. 110211 95.30757 1193. 10010
397. 118179 19.39003 83.02579 o. n 0.29805
0.02536 0.0 1091. 581511 120.685116 0.0
1212.26685 l1l1.67123 50.61634 l1.69239 2. 552115
80.61006 16.83751 o.o 0.01581 0.00955
0.0 -637.99194 0.00323 0.00036 o.o
0.00359
7.90000 322.01587 10. 41889 96. 76445 493. 10010
IIQII, 70361 17.92668 77. 1'.>945 0.0 0.26408
0.02072 0.0 1551, ll8364 158.54175 0,0

1110.02539 39.88182 56.88261 3.23552 2.27879


82.073111 15.64784 0.0 0.01581 0.00892
0.0 -591. 27002 0.00291 0.00030 0.0
0.00320
10.00000 322.01587 8.79963 97. 26727 1193, 10010
408.83057 17.08974 73.87587 0.0 0.24395
0.01826 0,0 1823.59814 178.52635 0.0
2002.12427 36.88847 60.37885 2.73267 2. 10775
82.91035 111.98193 0.0 0.01581 0.00857
0.0 -564.80444 .0.00270 0.00026 0.0
0.00296
15.70000 322.01587 6. 651911 97. 931,22 493. 10010 I-'
417.20581 15,39125 67.28900 0.0 0.20212 V,

0.01367 0.0 2388.41602 214.10533 o.o (X)

2602. 521211 30. 511176 67. 392119 2.06572 1.74511


811.60881 13. 611612 0.0 0.01581 0.00786
0.0 -511.36841 0,00225 0.00020 0.0
0. 0021,5
16.50000 322.01587 6.47256 97. 989911 493. 10010
418. 144011 15.20098 66.55524 0.0 o. 19737
0.01320 o.o 2452.52808 217. 63335 0.0
2670.16138 29.81627 68.17369 2.01001 1.70366
84. 79901, 13.49731 0.0 0.01581 0.00778
0.0 -505.39746 0.00220 0.00020 0.0
0.00239
20.00000 322.01587 5.86034 98. 18010 493. 10010
1121.75708 111,116826 63. 731,39 0.0 0. 17897
0. 011112 0.0 2700.85815 230.30612 0.0
2931.161!06 27.00284 71.17723 1.81989 1.54290
85.53188 12.92526 0.0 0.01581 0.00748
0,0 -482.111797 0.00200 0. 00017 0.0
0. 00217
:
PROC GPLOT DATA=FULLGOUT=PS1COD;
PLCH SlCOD*TltETAC;
LABEL S1COD==
11 :

SYMBOL1 V=SqlJARE I =SPLINE.;


ITfl[l .F:=SIMPLEX .H=l .A"'90 EFFLUENTCOD (MG/L);
PROCGPI.OT DATA=FULI.GOUT=PECODR;
PLOT ECODR*TllrT AC;
l.Al3[1. [COOR='I;
SYMROLlV=SQUAREl=SPLINE;
TITLEl .FcccSIMPLEX.11=1 .A=90 'COD REMOVAL PERCENTAGE';
PROCGPI.Of DATA=FULLGOUT=PEFFNIT;
PLOT EFFNIT*THETAC;
LABELHFNIT=' I;
SYM130L1V=SQUAREl=JOIN;
TITLEl .F=SIMPLEX .11=1 .A=90 'EFFLUENT Nll/1-N (MG/L)';
PROCGPI.OT DATA=FULLGOUT=PENH4R;
PLOT ENIIIIR*THE.TA;
I ABEi. ENIIL1R='';
SYMOOLlV=SQUAREl=JOIN;
TITIF:1 .F=SIMPLEX .H=l .A=90 'NH4-N REMOVAL PERCENTAGE';
PROCGPI.OT DAIA=FULLGOUT=PENITC;
PLOT(EFFNIT EFN02 EFN03)*THETAC/ OVERLAY;
...,
V,
\C
LABEL EFfNI f=' ';
SYMBOLlV=SQUAREl=JOIN;
SYMBOL2V=TRIANGLEl=JOIN;
SYMBOL3V=DIAMOND l=,JOIN;
llTLEl .f=SIMPLEX .11:=1 .A=90 EFFLUENTNllHOGEN (MG/L);
FOOTNOJ[l .f=SIMPLEX .H=l .. J=C 1 --EFF Nfl/1-N'·
f001NOIE2 .F=SIMPLEX .11=1 .J=C 1 --EFF N02-N':
F001NOTE3 .F=SIMPLEX .11=1 .J=C ' --EFF N03-N';
PROCGPLOT DATA=FULLGOUT=PYOCN;
PLOT(YOCOD YONITS YONITB)·H-lllETAC / OVERLAY;
LABELYOCOD= I I YONI TS=I I YONI TB=I I;
SYMBOi1 V=SQUARLI =SPLINE.;
SYMBOl2V=TRIANGI.E!=JOIN;
SYMBOL3V=DIAMOND !=JOIN;
TIILEl .F=SIMl'LEX .H=l .A=90' OBSERVEDYIELD COE.FF.';
roornorEl .F=SIMPLEX .11=1 .J=C I COD HEMOVAL';
l'OOINOTE2 .F=SIMPI.EX .11=1 .J=C' -- NIil! REMOVAL'·
f001NOlE3 .F=SIMPLEX .H=l .J=C ' -- N02 REMOVAL';
PROCGPLOT DATA=fULLGOUT=PXD;
l'I.OT(XCODXNITS XNITB XMI.SS)*THETAC / OVERLAY;
LABELXCOD='I XNITS=' I XNITB=' I XMLSS='';
SYMBOi 1 V=SQUAR[!~SPLINE;
SYMUOL.2 V=TRIANGLEl=JOIN;
SYMBOL3V=DIAMOND !=JOIN;
SYMBOLII V=STAR l=JOIN;
TITLE1 ,f,=SIMPLEX .11=1 .A=90 VSS CONCENTRATION (MG/L);
FOOTNOfEl .F=SIMPLEX .H=l .J=C 1 HETEROTROPHSI ;

FOOTNOfE2.F=SIMPLEX .H=l .J=C 1 N11ROSOMONADS1 ;

FOOTNOTE3.F=SIMPLEX .11=1 ,J=C 1 NITRODACTERS1 ;

FOOTNOT[4.F=SIMPLEX .H=l ,J=C ' TOT. vss CONG.I;


PROCGPLOT DATA=FULLGOUT=PCO;
l'LOT(C5C C02C C6C)~·TIIETAC/ OVERLAY;
LADELC5C=11 C02C=11 C6C='';
SYMBOLlV=SQUAREl=JOIN;
SYMDOL2V=TRIANGLEl=JOIN;
SYMROL3V=DIAMONDl=JOIN;
Till.El .F=SIMPLEX ,11=1 ,A=90' CARBONDISTRIBUTION PERCENTAGE';
FOOTNOH1 .F=SIMPLEX .11=1 .J=C I SYN-C';
F001NOTE2 .F=SIMPLEX .H=l ,J=C ' co2-c 1 •
FOOTNOTf3.F=SIMPLEX ,11=1 .J=C 1 -- GLU-c'i
FOOTNOT[II;
PROCGPLOT DATA=FULLGOUT=PND;
PLOT(C5N NH4N N02N N03N)*THETAC/ OVERLAY;
LABELC5N='' Nlll1N='' N02N=' 1 N03N=' '; .....
SYMDOLlV=SQUARE!=SPLINE; a,
SYMDOL2V=TRIANGLEl=JOIN; 0
SYMROL3V=DIAMOND(:JOIN;
SYMBOL4V=STAR l=JOIN;
TITIEl .F=SIMPL.EX .H=l .A=90 'NITROGENDISTRIBUTION PERCENTAGE';
roorNOTEl .F=SIMPL.E.X.llc=l .J=C I SYN-N';
F001NOTE2 ,F=SIMPLEX .11=1 .J=C I NHl1-N';
FOOINOTE3.F=SIMPLEX .ll=l .J=C' -- N02-N';
FOO'fNOT[II.F=SIMPLEX .11=1 .J=C 1 -- N03-N';
PROCGPLOT DAlA=FULLGOUT=P02RD;
PLOT(02REQD02RQNS 02RQND)*THETAC / OVERLAY;

l.AIH.I. 02RE(~I):.:'' 02HQNS='' 02RQNB=1 1 ;


SYMBOl.1V:.SQUAR[ l=JOIN;
SYMUOl.2V=TRIANGLEl=JOIN;
SYMBOL3V=DIAMOND l=JOIN;
SYMBOL 11;
TITL[1 .F=SIMPL[X ,11=1 ,A:=90 1 OXYGENREQIJI REMENTS(LBS/DAY)1 ;
F001N01E1 .F=SIMPLEX .11=1 .J=C I TOT-REQD,1 ;
FOOJNOTE2.F=SIMPLEX .ll-'l .J=C Nll4-0X ID.';
fOOTN01E3 .F=SIMPLEX .11=1 .J=C N02-0X ID. 1 ;
PHOCGPLOT DATA=FULLGOUT=PALKCHG;
PLOT Al.KCIIG*HIET AC;
L/\BLL Al l<CIIGcc1 1 ;
SYMl\01.1 V-=SQIIARE I-c-,JOIN;
TITLl:1 ,[:-:SIMPLEX .11=1 .A=90 'ALKALINITY CIIANGF (MG/LAS CAC03)';
F001N01F1;
PROC GPLOI IJAIA=rtJLL GOUl=PSLUDGE;
PLOT(CODSP NITSSP NITBSP TOTSP)*THETAC / OVERLAY;
IABEI. CODSP=' 1 NI TSSP 7 '' N 11 BSP='' TOTSP=' ';
SYMBOl.1 VccSQU/\RE !=SPLINE;
SYMBOL2 V=TRIANGLE l=JOIN;
SYMBOL3 V~TRIANGLE l=JOIN;
SYMIIOLII V=Sl/\R !=JOIN;
TITLEl .F=SIMPLEX .H=1 .A=90 'SLUDGE PRODUCFION \LBS/DAY)';
FOOTNOT[l .F=SIMPL[X .11°1 .J=C 1 COD REMOVAL ;
IOOINOl[2 .F=SIMPLEX .11=1 .J~C 1 NH4 REMOVAL';
FOOINOT[3 .F=SIMPLEX .11=1 .J=C 1 N02 REMOVAL 1 ;
FOOINOIEl1 .F=SIMPL[X .11=1 .J=C ' TOT REMOVAL' ;
DATA ALL;
SET PSlCOD PECODR PEFFNIT PENH4R PENITC PYOCN PXD PCD PND P02RD
PALl<CIIG PSLUDGE;
PROC GREPLAY DATA=ALL; I-'
O'\
I-'
APPENDIXB

RAWDATAFOR EACHOF THE FIVE

STEADYSTATEDATACOLLECTION
PERIODS

162
Table I

Raw Data for aC = 3.9 Days and COD:TKN= 6.07:1

NO -N
3
COD TKN Concentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

10-1 333 14 95.8 53.5 21.5 59.8 10.6 19.1 -44.5 24.5

10-2 352 12 96.6 54.7 20.0 63.4 10.3 15.8 -34.8 25.9 I-'
a-
w
10-3 343 28 91.8 54.1 16.2 70.1 10.9 13.3 -18.0 29.3

10-4 347 18 94.8 54.7 12.3 77.5 10.6 9.7 8.5 31.7

10-5 338 16 95.3 55.3 9.6 82.6 10.9 8.5 22.0 32.8

10-6 343 12 96.5 56.5 10.2 81.9 11.2 8.8 21.4 33.8

10-7 348 45 87.1 55.9 16.3 70.8 12.1 13.3 -9.0 31.3

AVG. 343 21 93.9 55.0 15.1 72.5 10.9 12.6 -13.5 29.9
Table I (continued)

NO2 -N Biological
Concentration Alkalinity solids ec pH

Remov. Total · Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

10-1 7.2 386 348 9.8 975 1130 6.0 4.3 7.8/7.5/7.8

10-2 8.2 394 324 17.8 850 1090 9.5 4.5 7.9/7.6/7.8
I-'
10-3 7.7 399 308 22.8 900 1060 5.5 4.3 7.9/7.6/7.8 °'
10-4 7.5 395 279 29.4 835 985 7.5 4.2 7. 9/7 .4/7 .5

10-5 7.2 391 270 30.9 895 1020 6.0 4.2 7.9/7.5/7.7

10-6 6.2 392 265 32.4 770 845 18.0 3.5 7.9/7.5/7.5

10-7 3.0 389 173 55.5 670 680 54.0 2.3 7.9/7.4/7.3

AVG. 6.7 392 281 28.3 842 973 15.2 3.9 7 .9/7 .5/7 .6
Table I (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effll Effl.
Date (oC) (mg/1) (mg/1/hr) (hr- 1) (hr-) (mg/1)

10-1 20.5 4.95 28.50 5.64 24.27 10.00

10-2 20.0 4.05 30.86 4.12 27.73 11.54

10-3 20.0 3.45 26.67 4.43 27.65 9.47 I-'


°'
V,

10-4 20.0 2.95 29.33 4.42 27.26 9.58

10-5 20.0 2.85 28.00 4.43 28.38 9.17

10-6 20.0 3.00 25.50 4.11 27 .72 9.20

10-7 20.5 4.00 21.00 4.17 24.73 9.04

AVG. 20.0 3.60 27.12 4.44 26.82 9.71


Table II

Raw Data for 8 = 4.2 Days and COD:TKN= 6.07:1


C

NO -N
3
COD TKNConcentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

11-23 344 29 91.6 55.0 7.6 86.2 10.1 5.0 50.5 59.5

11-24 337 19 94.4 53.8 5.1 90.5 11.1 4.5 59.5 54.0 I-'
°'
°'
11-25 351 23 93.4 54.4 7.0 87.1 10.6 4.5 57.5 54.0

11-26 343 21 93.9 54.4 8.8 83.8 11.1 4.5 59.5 35.5

11-27 334 15 95.5 53.5 7.3 86.4 10.6 5.5 48.1 29.8

11-28 348 15 95.7 55.0 9.4 82.9 10.1 7.1 29.7 31.8

11-29 362 17 95.3 55.3 10.1 19.8

AVG. 346 20 94.2 54.5 7.5 86.2 10.5 5.2 50.8 40.6
Table II (continued)

NO -N Biological
2 ec
Concentration Alkalinity solids pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

11-23 2 .1 388 245 1050 1140 76.0 2.9 7.9/7.6/7.5

11-24 1.4 385 245 1030 1120 42.0 3.7 8.0/7.5/7.5


...,
0\
11-25 1.1 383 246 1020 1120 32.0 4.1 8.0/7.5/7.6 -...J

11-26 0.8 392 273 1050 1120 14.0 4.7 8.0/7.5/7.7

11-27 0.6 372 253 1100 1110 9.5 4.6 8.0/7 .5/7.7

11-28 0.8 385 281 1220 1170 8.0 4.5 8.0/7.5/7.8

11-29 1.0 396 355 1360 1340 4.0 4.8 7.9/7.6/7.9

AVG. 1.1 386 271 1119 1160 26.5 4.2 8.0/7 .5/7 .7
Table II (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquyr Effli Effl.
Date (oC) (mg/1) (mg/1/hr) (hr-) (hr-) (mg/1)

11-23 21.0 4.05 30.00 5.50 26.00 9.50

11-24 21.0 2.95 33.33 6.25 21.88 8.28

11-25 20.5 a.so 37 .71 4.70 28.99 8.83


I-'
Q'\
11-26 20.0 0.85 29.00 3.39 26.17 9.41 00

11-27 20.5 2.05 33.60 4.58 26.22 9.38


'
11-28 20.5 1.95 33.00 4.20 25.76 9.81

11-29 20.5 38.00 26.06 11.00

AVG. 20.5 2.10 33.52 4.55 25.87 9.46


Table III

Raw Data for 0C = 8.5 Days and COD:TKN= 6.07:1

NO -N
COD TKNConcentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

8-17 325 19 94.2 68.1 4.3 93.7 11.0 0.9 91.8 39.5

8-18 316 25 92.1 55.2 3.4 93.8 17.5 0.3 98.3 49.0 I-'
a,
\0

8-19 305 23 92.5 54.6 3.4 93.8 11.0 0.3 97.3 51.0

8-20 321 28 91.3 57 .9 4.0 93.1 11.3 0.6 94.7 48.0

8-21 325 26 92.0 56.7 5.8 89.8 15.9 0.3 98.1 46.0

8-22 323 19 94.1 57.3 5.8 89.9 12.3 0.6 95.1 41.5

8-23 330 30 90.9 57.3 4.3 92.5 12.6 0.3 97.6 41.0

AVG. 321 24 92.5 58.2 4.4 92.4 13.1 0.5 96.2 45.1
Table III (continued)

NO -N Biological
Concen~ration Alkalinity solids ec pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/ 1) (days) Inf/ML/Effl.

8-17 6.2 373 173 53.6 1385 1675 0.5 9.3 7.7/7.2/7.4

8-18 4.0 374 177 52.7 1390 1640 1.0 9.0 7.7/7.2/7.5

8-19 1.8 382 186 51.3 1480 1650 1.5 8.5 7.4/7.1/7.2 1--'
0
8-20 1.1 379 197 48.0 1450 1640 5.5 8.2 7.8/7.2/7.5

8-21 0.8 365 183 49.9 1455 1640 4.0 8.4 7.8/7.2/7.5

8-22 1.0 372 192 48.4 1500 1715 10.0 7.9 7.7/7.3/7.5

8-23 0.7 372 197 47.0 1525 1780 11.5 8.0 7.7/7.2/7.4

AVG. 2.2 374 186 50.1 1455 1677 4.9 8.5 7.7/7.2/7.4
Table III (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effll Effl.
Date (oC) (mg/1) (mg/1/hr) (hr- 1) (hr-) (mg/1)

8-17 20.0 4.50 21.60 4.55 26.07 9.25

8-18 20.0 3.80 18.67 3.85 26.70 8.65

8-19 20.0 3.60 21.33 3.35 26.00 9.97


1--'
8-20 20.5 3.85 20.31 3.06 28.19 10.49 -...J
1--'

8-21 20.0 3.95 17.45 3.14 26.25 9.50

8-22 20.5 3.75 18.46 3.59 29 .07 8.89

8-23 20.0 3.85 17.00 3.15 28.58 9.25

AVG. 20.0 3.90 19.26 3.48 27.27 9.43


Table IV

Raw Data for aC = 13.8 Days and COD:TKN= 6.07:1

NO -N
COD TKNConcentration NH3-N Concentration Concen~ration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

12-20 332 20 94.0 56.0 4.8 91.4 10.0 2.5 75.0 42.5

12-21 322 19 94.1 54.5 4.8 91.2 9.5 2.0 78.9 45.0 I-'
-...J
N
12-22 324 17 94.8 54.8 4.8 91.2 10.0 2.5 75.0 42.5

12-23 322 19 94.1 54.8 3.9 92.9 10.0 2.0 80.0 42.5

12-24 324 17 94.8 54.5 3.3 93.9 9.5 2.5 73.7 41.5

12-25 324 17 94.8 56.3 3.6 93.6 9.5 2.5 73.7 42.5

12-26 320 20 93.8 54.2 3.9 92.8 10.0 2.0 80.0 45.0

AVG. 324 18 94.4 55.0 4.2 92.4 9.8 2.3 76.5 43.0
Table IV (continued)

NO -N Biological
Concentration Alkalinity solids ec pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf /ML/Ef fl.

12-20 0.4 366 185 49.5 2005 2305 9.5 13.3 7.9/7.4/7.6

12-21 0.4 366 185 49.5 1870 2290 7.5 14.4 7.9/7.2/7.6

12-22 0.4 367 187 49.0 1980 2370 7.5 14.2 7.9/7.3/7.5 .....
-.J
w
12-23 0.4 365 187 48.8 2050 2180 6.0 12.9 7.9/7.4/7.6

12-24 0.4 365 185 49.3 2140 2340 7.0 13.1 7.9/7.4/7.7

12-25 0.5 363 183 49.6 2150 2410 5.5 13.7 8.0/7.5/7.7

12-26 0.6 374 185 50.5 2030 2420 4.5 14.8 7.9/7.4/7.6

AVG. 0.4 367 185 49.5 2032 2331 6.8 13.8 7.9/7.4/7.6
Table IV (continued)

Temperature DO R Ka DOSAT
Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effli Effl.
Date (oC) (mg/1) (mg/1/hr) (hr- 1) (hr-) (mg/1)

12-20 20.5 3.70 39.10 5.36 27.66 10.99

12-21 21.0 3.70 41.25 6.54 27.51 10.01


12-22 21.0 3.20 43.38 6. 1.0 27.24 10.31
....
.....
12-23 21.0 3.95 39.16 6.44 26.08 10.03

12-24 20.0 4.05 43.38 7.10 26.59 10.16

12-25 20.5 2.60 38.40 4.56 27.48 11.02

12-26 20.5 3.45 44.00 6.04 27.85 10.73

AVG. 20.5 3.50 41.24 5.93 27.20 10.46


Table V

Raw Data for eC = 21.0 Days and COD:TKN= 6.07:1

NO -N
COD TKN Concentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

7-24 339 18 94.7 53.8 o.o 100.0 9.4 o.o 100.0 53.0

7-25 339 16 95.3 55.0 o.o 100.0 0.3 52.0 f--'


-..J
V,

7-26 336 16 95.2 55.6 o.o 100.0 0.3 40.5

7-27 338 16 95.3 52.3 o.o 100.0 10.6 o.o 100.0 41.5

7-28 332 18 94.6 50.6 0.6 98.8 10.9 0.0 100.0 40.5

7-29 336 18 94.6 50.6 o.o 100.0 12.0 1.5 87.5 64.0

7-30 330 24 92.7 46.9 o.o 100.0 7.6 0.3 96.1 64.0

AVG. 336 18 94.6 52.1 0.09 99.8 10.1 0.3 97.0 50.8
Table V (continued)

NO -N Biological
Concenrration Alkalinity solids ec pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

7-24 374 179 52.1 1876 2168 5.0 20.1 7.6/7.4/7.4

7-25 390 202 48.2 2080 2116 4.0 18.1 7.8/7.4/7.5

7-26 392 206 47.4 2064 2464 3.5 21.6 7.7/7.5/7.5 .....
....,
7-27 374 195 47.9 2196 2528 0.5 22.1 7.6/7.4/7.5 °'
7-28 376 195 48 .1 2128 2548 0.5 23.0 7 .8/7 .3/7 .5

7-29 385 200 48.1 2196 2588 7.0 20.1 7.7/7.2/7.4

7-30 387 194 49.9 2288 2612 o.o 22.2 7.7/7.3/7.5

AVG. 383 196 48.8 2118 2432 2.9 21.0 7.7/7.4/7.5


Table V (continued)

Temperature DO. R
~- DOSAT
Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effll Effl.
Date (oC) (mg/1) (mg/1/hr) (hr- 1) (hr-) (mg/1)

7-24 20.0 4.90 25.50 6.82 25.87 8.64

7-25 19.5 4.40 20.94 4.12 26.17 9.48

7-26 19.5 3.40 27.00 4.86 23.15 8.95


I-'
"'-I
7-27 20.5 3.60 48.00 8.66 25.76 9.14 "'-I

7-28 19.0 3.20 45.45 7.70 29.59 9.10

7-29 19.0 2.80 46.07 6.88 27.87 9.50

7-30 19.5 3.15 32.73 5.59 25.77 9.01

AVG. 19.5 3.65 35.10 6.42 26.31 9.12


Table VI

Raw Data for eC = 2.2 Days and COD:TKN= 0.65:1

NO -N
COD TKN Concentration NH.3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

9-30 340 39 88.5 498.2 449.1 9.9 461.8 442.0 4.3 3.3

10-1 345 71 79.4 501.3 453.6 9.5 472.4 446.6 5.5 3.1 ....
"'
00

10-2 331 85 74.3 498.2 455.9 8.5 458.7 442.0 3.6 2.6

10-3 340 75 77 .9 498.2 447.6 10.2 460.3 446.6 3.0 3.0

10-4 333 51 84.7 496.7 452.1 9.0 461.8 439.0 4.9 3.6

10-5 340 81 76.2 501.3 447.6 10.7 457.2 439.0 4.0 4.7

10-6 343 90 73.8 498.2 447.6 10.2 461.8 437.5 5.3 6.6

AVG. 339 70 79.4 498.9 450.5 9.7 462.0 441.8 4.4 3.8
Table VI (continued)

NO -N Biological
Conceniration Alkalinity solids ec pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

9-30 23.2 3570 3396 4.9 495 575 40.0 2.7 8.0/8.5/8.6

10-1 20.3 3629 3401 6.3 405 480 43.0 2.4 8.0/8.5/8.6
....
-.J
10-2 19.3 3559 3467 2.6 445 565 43.5 2.7 8.0/8.5/8.6 1.0

10-3 21.0 3575 3407 4.7 415 455 48.5 2.1 8.0/8.6/8.7

10-4 24.3 3548 3418 3.7 515 415 58.0 1.6 8.0/8.5/8.6

10-5 26.3 3510 3368 4.0 360 440 57.5 2.0 8.0/8.6/8.6

10-6 30.0 3505 3343 4.6 355 420 58~5 2.0 7.9/8.5/8.5

AVG. 23.5 3557 3400 4.4 427 479 49.9 2.2 8.0/8.5/8.6
Table VI (continued)

Temperature DO R Ka DOSAT
Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effl. Effl.
Date (OC) (mg/1) (mg/1/hr) (hr- 1) (hr-l) (mg/1)

9-30 20.0 5.95 26.31 6.61 42.97 9.93

10-1 20.0 6.30 27.00 7.87 49.35 9.73

10-2 20.0 5.65 24.50 10.70 46.58 7.94


>-'
co
10-3 20.0 5.35 18.30 3.87 45.41 10.08 0

10-4 20.0 5.25 15.90 4.02 45.96 9.21

10-5 20.0 5.50 19.50 5 .13 45.34 9.30

10-6 20.0 5.55 18.30 4.88 45.34 9.30

AVG. 20.0 5.65 21.40 5.77 45.85 9.36


Table VII

Raw Data for eC = 4.3 Days and COD:TKN= 0.65:1

NO -N
COD TKN Concentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

11-23 325 96 70.5 502.0 332.6 33.7 456.6 332.6 27.2 6.0

11-24 330 60 81.8 491.4 335.7 31.7 446.0 328.1 26.4 18.0 ....
....
CXl

11-25 323 49 84.8 491.4 326.6 33.5 452.1 320.5 29.1 20.3

11-26 351 38 89.2 502.0 316.0 37.1 455.1 326.6 28.2 25.0

11-27 344 38 89.0 500.5 323.6 35.3 455.1 322.1 29.2 16.0

11-28 371 42 88.7 500.5 337.2 32.6 452.1 337.2 25.4 10.0

11-29 348 62 82.8 503.5 347.8 30.9 449.1 349.3 22.2 7.8

AVG. 342 55 83.9 498.8 331.4 33.6 452.3 330.9 26.8 14.4
Table VII (continued)

NO -N Biological
Concentration Alkalinity solids eC pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

11-23 137.4 3711 2738 26.2 860 1130 28.0 5.0 7.9/8.2/8.3

11-24 145.3 3471 2809 19.1 1060 1090 29.0 4.0 8.0/8.3/8.4

11-25 135.0 3597 2666 25.9 810 1015 26.0 4.7 8.0/8.2/8.3 f...l
00
N
11-26 143.9 3597 2600 27.7 940 1105 29.0 4.5 8.0/8.2/8.5

11-27 138.0 3652 2573 29.3 885 1085 33.5 4.4 8.0/8.2/8.3

11-28 120.2 3559 2738 23.1 935 985 39.5 3.7 8.0/8.3/8.4

11-29 117 .2 3542 2803 20.9 920 1015 34.5 3.9 7.9/8.2/8.3

AVG. 133.9 3590 2704 24.7 916 1061 31.4 4.3 8.0/8.2/8.4
Table VII (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effl. Effl.
Date (OC) (mg/1) (mg/1/hr) (hr- 1) (hr- 1) (mg/1)

11-23 20.0 2.45 111.00 20.71 44.57 7.81

11-24 20.0 2.50 81.00 12.82 44.99 8.82

11-25 20.0 3.05 104.00 19.85 40.32 8.29


.....
00
11-26 19.5 2.90 106.00 16.21 42.29 9.44 vJ

11-27 20.0 3.35 108.00 17.12 43.64 9.66

11-28 20.0 3.15 100.00 16.29 44.77 9.29

11-29 20.0 2.95 94.28 12.34 46.49 10.59

AVG. 20.0 2.90 100.61 16.15 43.87 9.13


Table VIII

Raw Data for eC = 7.9 Days and COD:TKN= 0.65:1

NO -N
COD TKNConcentration NH3-N Concentration Concentration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

8-17 271 32 88.2 489.0 230.0 53.0 463.6 237.6 48.7 19.8

8-18 290 34 88.3 492 .1 274.4 44.2 465.6 270.6 42.7 16.8 I-'
00
.t-
8-19 295 40 86.4 495.2 256.0 48.3 462.5 263.7 43.0 14.8

8-20 290 44 84.8 481.4 266.7 44.6 466.3 265.2 43.1 19.3

8-21 286 48 83.2 489.0 260.6 46.7 454.9 260.6 42.7

8-22 307 38 87.6 487.5 272 .9 44.0 461.0 263.7 42.8 19.3

8-23 297 48 83.8 487.5 282.1 42.1 460.3 269.8 41.4 11.3

AVG. 291 41 85.9 488.8 263.2 46.2 462.0 261.6 43.4 16.9
Table VIII (continued)

NO -N Biological
Concen~ration Alkalinity solids ec pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

8-17 210 3397 1944 42.8 2115 2555 20.5 8.0 7.8/8.0/8.1

8-18 184 3537 2176 38.5 2120 2630 21.5 8.2 7.8/8.0/8.2

8-19 186 3591 2245 37.5 2155 2575 23.0 7.8 7.5/8.0/8.1 1--
00
V,

8-20 179 3624 2267 37.4 2070 2515 20.5 8.1 7.9/7.9/8.2

8-21 181 3570 2201 38.3 2050 2455 22.5 7.8 7.9/8.1/8.3

8-22 174 3526 2267 35.7 2060 2525 22.5 8.0 7.8/8.2/8.2

8-23 156 3553 2267 36.2 2110 2490 26.0 7.6 7.8/8.0/8.2

AVG. 181 3543 2195 38.0 2097 2535 22.4 7.9 7.8/8.0/8.2
Table VIII (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effli Effl.
Date (OC) (mg/1) (mg/1/hr) (hr- 1) (hr-) (mg/1)

8-17 20.0 1.15 112.80 12.33 44.37 10.30

8-18 20.0 0.90 132.00 13.46 46.90 10.71

8-19 20.0 0.85 180.00 19.59 49.83 10.04


.....
00
8-20 20.0 0.95 168.00 16.99 44.81 10.84 Q\

8-21 20.0 0.75 172.00 19.22 45.69 9.70

8-22 20.0 0.70 102.00 13.78 43.84 8 .10

8-23 20.0 0.60 180.00 19.29 37.79 9.93

AVG. 20.0 0.85 149.54 16.43 44.75 9.95


Table IX

Raw Data fore = 15.7 Days and COD:TKN= 0.65:1


C

NO -N
COD TKN Concentration NH3-N Concentration Conceniration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

12-22 338 11 96.7 498.2 254.3 49.0 457.5 248.3 45.7 42.7

12-23 304 21 93.1 499.7 270.9 45.8 465.0 261.9 43.7 43.0 ....
00
--.J

12-24 336 19 94.3 493.6 270.9 45.1 456.0 264.9 41.9 41.0

12-25 331 24 92.7 493.6 273.9 44.5 456.0 267.9 41.3 42.0

12-26 336 34 89.9 468.1 276.9 40.8 276.9 42.3

12-27 342 30 91.2 230.3 230.3 42.3

12-28 342 30 91.2 493.6 251.3 4·9.1 459.0 248.3 45.9 42.5

AVG. 333 24 92.8 491.1 261.2 46.8 458.7 256.9 44.0 42.3
Table IX (continued)

NO -N Biological
Conceniration Alkalinity solids ac pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

12-22 154.8 3594 2061 42.7 3045 3700 20.5 20.3 8.0/8.1/8.3

12-23 181.5 3530 2194 37.8 3270 2390 27.5 11.5 8.0/8.1/8.3

12-24 180.5 3387 2258 33.3 3290 3040 23.0 15.3 8.0/8.2/8.3 I-'
00
00
12-25 174.9 3530 2226 36.9 3590 3710 25.5 17.0 8.1/8.2/8.4

12-26 179.8 3493 2258 35.4 3480 3240 28.5 14.7 8.0/8.2/8.3

12-27 161.6 3616 2338 35.3 3510 3400 20.0 16.8 8.0/8.2/8.4

12-28 175.6 3499 2381 32.0 3350 3100 30.0 14.3 8.0/8.2/8.3

AVG. 172.6 3521 2245 36.2 3362 3226 25.0 15.7 8.0/8.2/8.3
Table IX (continued)

Temperature DO R KLa DOSAT


Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liquor Effl. Effl.
Date (oC) (mg/1) (mg/1/hr) (hr- 1) (hr- 1) (mg/1)

12-22 20.5 0.85 150.00 16.16 39.28 10.13

12-23 20.5 1.65 126.00 15.05 35.86 10.02

12-24 19.5 0.90 162.00 15.23 42.29 11.54


. I-'
00
12-25 20.0 0.80 180.00 17.66 40.57 10.99 1.0

12-26 19.5 0.80 180.00 17.98 35.33 10.81

12-27 19.5 0.75 168.00 22.58 39.24 8.19

12-28 19.5 0.65 120.00 16.42 39.51 7.96

AVG. 20.0 0.90 155.14 17.14 38.87 9.95


Table X

Raw Data for eC = 16.5 Days and COD:TKN= 0.65:1

NO -N
COD TKNConcentration NH3-N Concentration Conceniration

Remov. Remov. Remov.


Feed Effl. Effie. Feed Effl. Effie. Feed Effl. Effie. Effl.
Date (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1) (mg/1) (%) (mg/1)

7-28 270 36 86.7 499.8 156.4 68.7 450.6 162.5 63.9 44.3

7-29 272 44 83.8 475.2 156.4 67.1 470.2 166.3 64.6 31.0 t-"
\0
0
7-30 269 53 80.3 489.9 159.4 67.5 470.2 170 .1 63.8 24.5

7-31 343 48 86.0 129.0 471.7 161.8 65.7 25.5

8-1 301 10 96.7 148.2 468.7 152.7 67.4 41.8

8-2 335 11 96.7 486.9 468.7 152.7 67.4

8-3 337 19 94.4 490.6 189.0 61.5 468.7 163.3 65.2 18.8

AVG. 304 32 89.5 488.5 156.4 68.0 467.0 161.3 65.5 31.0
Table X (continued)

NO -N Biological
Concentration Alkalinity solids Sc pH

Remov. Total Aeration Total


Effl. Feed Effl. Effie. System Basin Effl. System
Date (mg/1) (mg/1) (mg/1) (%) (mg/1) (mg/1) (mg/1) (days) Inf/ML/Effl.

7-28 3450 1632 52.7 2696 3272 25.0 17.2 7.8/7.9/7.9

7-29 3608 1779 50.7 2716 3396 23.5 17.7 7.9/7.9/8.1

7-30 3624 1747 51.8 2904 3560 28.0 16.7 7.9/7.9/8.1 I-'
I.O
I-'
7-31 3526 1697 51.9 3192 3720 41.0 14.4 7.8/7.9/8.0

8-1 3581 1648 54.0 2940 3632 30.5 16.8 7.8/7.9/8.0

8-2 3422 1643 52.0 2935 3825 34.0 17.0 7.8/7.8/7.9

8-3 3504 1674 52.2 3185 3690 29.5 16.0 7.7/7.8/8.0

AVG. 3531 1689 52.2 2938 3585 30.2 16.5 7.8/7.9/8.0


Table X (continued)

Temperature DO R Ka DOSAT
Mixed Mixed Mixed Mixed
Liquor Liquor Liquor Liqu~r Effll Effl.
Date (OC) (mg/1) (mg/1/hr) (hr-) (hr-) (mg/1)

7-28 20.0 0.40 192 .oo 21.33 44.72 9.40

7-29 20.0 0.40 240.00 27.00 44.92 9.29

7-30 20.0 0.35 240.00 27.71 35.70 9.01


I-'
I.O
7-31 20.0 0.35 234.00 24.32 40.61 9.97 N

8-1 20.0 0.30 252.00 27.04 46.05 9.62

8-2 20.0 0.35 288.00 27 .77 45.98 10. 72

8-3 20.0 0.35 288.00 30.77 47.67 9.71

AVG. 20.0 0.35 247.71 26.58 43.66 9.67


The two page vita has been
removed from the scanned
document. Page 1 of 2
The two page vita has been
removed from the scanned
document. Page 2 of 2
OXYGENTRANSFERSTUDIES IN THE

COMPLETELY
MIXED ACTIVATEDSLUDGEPROCESS

by

Richard Oliver Mines, Jr.

(ABSTRACT)

Utilization of the activated sludge process is widespread

although many of the mechanisms that make it work are still relatively

misunderstood. Recent studies have indicated that dual substrate

limitations may occur in the process. Several misconceptions in the

basic fundamentals regarding the rates and mechanisms involved in oxy-

gen transfer to wastewater systems also exist.

This research investigation examined the effects of the mean cell

residence time and wastewater stoichiometry on the operation of the

completely mixed activated sludge process under a dual substrate limi-

tation. At low mean cell residence times (e) the system was growth
C

limited with respect to carbon and at high mean cell residence times

the system was oxygen limited. Oxygen transfer studies were conducted

to ascertain the relationship between the steady state oxygen transfer

coefficient (KLa) and the oxygen uptake rate of the mixed liquor

(R).
The objectives of this research were accomplished by operating

two continuous flow bench scale activated sludge units at COD:TKN


ratios bf 6.07:1 and 0.65:1. Reactor-1 was operated at a

COD:TKN= 6.07:1 and was always growth limited with respect to organic

carbon while Reactor-2 was operated at a COD:TKN= 0.65:1 and was

carbon limited at low mean cell residence times and oxygen limtied at

high ae values. Mean cell residence time served as the primary con-

trol parameter during the laboratory studies and was varied form

approximately 2.5 to 21.0 days.

Theoretical studies were also conducted in which biokinetic and

stoichiometric equations were used to develop a model to simulate the

process operating under carbon and oxygen limitations. The model was

found to yield results that were similar to the actual experimental

data collected. Further refinement of the model by including inhibi-

tion functions would result in a model with better predictability.

Examination of the experimental data collected during the labora-

tory study revealed several interesting conclusions. Operation of the

activated sludge process at a low COD:TKNratio (0.65:1) and under an

oxygen limitation at high mean cell residence times can result in high

levels of free ammonia and nitrite that will lead to a deterioration

in effluent quality. Increased removal efficiencies for COD, TKN and

NH3-N can be achieved by operating the process at a high COD:TKNratio

(6.07:1). Steady state oxygen transfer coefficients determined in the

mixed liquor of the reactors indicated there was a direct relationship

to the oxygen uptake rate of the activated sludge (R). This observa-

tion is quite significant since standard aeration theory states that

KLa is constant for a given aeration device. Nonsteady state KLa


values determined on the effluent from each reactor indicated that KLa

was a constant. Alpha and beta coefficients determined from nonsteady

state tests on wastewater effluent from each reactor showed no trend

with the mean cell residence time.

You might also like