Вы находитесь на странице: 1из 166

Effects of Biomass Growth on Pressure Drop in Biofilters

Fernando Morgan-Sagastume

A thesis subrnitted in conformity with the requirements


for the degree of Master of Applied Science
Graduate Department of Chemicd Engineering and Applied Chemistry
University of Toronto

@ Copyright by Fernando Morgan-Sagastume, 1999

1*1

National Cibrary

Bibliothque nationale
du Canada

Aquisitions and
Bibliographie Services

Acquisitions et
services bibliographiques
395, rue Wellington
Ottawa ON K I A ON4

of Canada

395 Wellington Street


OttewaON K1AON4
Canada

Canada

Ycur hb Vorre mhimce

Our file Notre refrence

The author has granted a nonexclusive licence allowing the


National Library of Canada to
reproduce, loan, distribute or sel1
copies of this thesis in microfonn,
paper or electronic formats.

L'auteur a accord une licence non


exclusive permettant la
Bibliothque nationale du Canada de
reproduire, prter, distribuer ou
vendre des copies de cette thse sous
la fonne de rnicrofiche/film, de
reproduction sur papier ou sur format
lectronique.

The author retains ownership of the

L'auteur conserve la proprit du


droit d'auteur qui protge cette thse.
Ni la thse ni des extraits substantiels
de celle-ci ne doivent tre imprims
ou autrement reproduits sans son
autorisation.

copyright in this thesis. Neither the


thesis nor substantial extracts from it
may be printed or otherwise
reproduced without the author's
permission.

ABSTRACT
EFFECTS OF BIOMASS G R O W H O N PRESSURE DROP IN BIOFILTERS
Fernando Morgan-Sagastume, Degree of Master of Applied Science, 1999
Graduate Department of Chernical Engineering and Applied Chemistry
University of Toronto

The effects of microbial mass accumulation and distribution in biofilter air pressure
losses were experimentally studied. A mode1 was proposed to predict pressure drop as a
fwiction of biomass concentration in a biofilter.
Two benchscale biofiltea, one packed with inert porous pellets (Nova lnerta) and the
other with wood chips, were operated with excess nutrients under similar conditions to treat air
polluted with methanol (loading rates of 100 to 150 g methanol/m3 bed/h) and generate biomass.
Uneven biomass distribution, described by localised high and low biornass concentrations in the
biofilter beds, was the key factor afTecting increased pressure drops due to extreme local
porosity reduction by clogging. nie highest pressure drops in the beds (2600 Pa/m in the wood

chip biofilter and 550 Pa/m in the Nova Inert biofilter) were caused by thick layers of biomass

with high moisture content and high extracellular polymeric substance (EPS) concentration.
Pressure &op varied exponentially with the amount of biomass and bed moisture, and
cumulative methanol consumption. Pressure drop increased significantly at a cntical point in
methanol removal and biomass accumuiation. Six-fold higher pressure drops were measured in
the wood chip biofilter compared to the Nova lnertm biofilter due to more biomass growth and
bed compaction.

ACKNOWLEDGEMENTS
1 would like to express my most sincere gratitude to al1 the people who, by one way or

another, guided, supported, and enriched my academic and persona1 development throughout
this work. Especially, 1 would like to aknowledge:
Professors D. Grant Allen and Brent E. Sleep for their fine and carefbl supervision,
guidance and research encouragement.
Martha Miller and Valene Farrugia for their continual support and advice in m i n g the
biofilters, and Prof. B. Saville for his helpfl advice analysing the tracer studies.
The Biofiltration Consortium of the Pulp and Paper Centre: Aracruz Cellulose S. A.

(Brazil), Avenor Inc., Domtar Packaging, Georgia-Pacific Corp., Nippon Paper Industries Co.
Ltd (Japan), and Weyerhaeuser Company, and the Natural Sciences and Engineering Research
Council (NSERC) of Canada for their financial support.
Staff and faculty of the Pulp and Paper Centre for their enthusiasm in the students'
professional and academic development.

My lab group: Prof. Allen, Chandra, Yang, Madjid, Christina, Raju, Martha, Anja, Ba,
Jessica and Choy-Siong, and other friends, for al1 their advice, ftienship, and solidarity.

My parents and brothers for their unconditional love and encouragement.

iii

LIST OF CONTENTS

ACKNOWLEDGEMENTS
LIST OF

LIST OF FIGURES
LIST OF
1

II
a

111
O

IV
a

VIII

a m a o a a a a e m o m m m o e m m m m o m m a o m a m m o ~ a o o a a ~ m m m o a o o o o e a o o o a o a a a a a o o ~ ~ o ~

XII

T A B L E S m m a a m a a m m a m a m m m m m m ~ a ~ o ~ a m o m a a a a a a a o a a a e a o a a o o a a o a o o a o a o

INTRODUCTION ...........................................mmm..m.m............................................................
1

THEBIOFLTRATION
PROCESS............................................................................................
1
1.2 STATEMENT
OF RESEARCH.................................................................................................
1
1.2.1
Problem and Hypothesis ............................................................................................
1
1.2.2
Objectives ................................................................................................................ -7
7
1.2.3
Signij'kance ................................................................................................................
1.3
THESI
s OUTLINE
................................................................................................................
3
1.1

THEORETICAL BACKGROUND AND LITERATURE REVIEW ~~~~~~mm~..m..m...~


4
BIOFILTRATION
AND ITS APPLKATION
TO THE TREATMENT
OF PULP.PAPER.AND
WOODAIRWASTEEMISSIONS
.......................................................................................4
The Process of BioJiltration.......................................................................................4
Application to the Pulp. Paper. and Wood Industries ............................................. 5
PARAMETERS
AFFECTING
PRESSURE
DROPIN B~OFILTERS
.................................................7
Superflcial Gas Velocity............................................................................................ 8
Bed Characteristics .................................................................................................10
Particle Size Distribution and Bed Depth ....................................................................................................
10
Nature and Composition of the Filter Medium .............................................................................................. I I
Moisture Content........................................................................................................................................12

Microbial Mass Growth........................................................................................... 14


BIOMASS
ACCUMULATION
A N D PRESSURE
DROPM FILTERPACKEDBEDS...................... 14
Biomass Development in Biofilters..........................................................................14
Factors Affecting Biomass Growth in a Filter Packd Bed .................................... 16
Support Material ............................................................................................................................................
16

Substrate Supply ............................................................................................................................................17

Nutritnts ........................................................................................................................................................t7
Other Factors .................................................................................................................................................19

Biomass Distribution in the Packed Bed of a Filter ................................................20


Effects of Biomass Growth on the Flow Characteristics......................................... 22
24
Biomass Growth Quantifcation in Packed B e h.....................................................
MODELLING
OF BIOMASS
GROWTH
AND ITS EFFECTS
ON PRESSURE
DROP......................26
26
Modelling Biomass Growth in Biofilters .................................................................
Modelling the Eflects of Biomass Accumulation on Pressure Drop........................29
Models Developed for Groundwater Flow in Porous Media ........................................................................ 29
Models Developcd for Biorcactors............................................................................................................

35

EXPERIMENTAL PROCEDURE ..................................................................................39


GENERAL
APPROACH
........................................................................................................
39
EXPERIMENTAL
SET-UP....................................................................................................
39
.......................................................................................................43
PACKING
MATERIALS
Prepration and Inoculation ...................................................................................
45
...........................................46
BEDSAMPLING
AND PRESSURE
DROP(hP) MEASUREMENTS
TRACER
STUDIES
..............................................................................................................
47
ANALYTICAL
TECHNIQUES
...............................................................................................48
....................*.. 48
Methanol and a-pinene Concentration Measurements ............
.
Bed Moisture Content and pH ................................................................................49
Interparticle Porosity...............................................................................................
49
Biomass Quantifiation ...........................................................................................
49
Biomass Detachment Technique ...................................................................................................................
50
Detached Biolaycr ........................................................................................................................................ 50

Total and Volatile Solids (TS & VS) .............................................................................................................50


Chemical Oxygcn Dcmand (COD) ................................................................................................................ 50
52
Total Protein ..................................................................................................................................................

4.1.1

.........................................................................................54
BIOFILTRATION
OF METHANOL
Performance ............................................................................................................
54

4.1.1.1

Ovcrall Methanol Removal........................................................................................................................... 5 4

4.1.1.2

Axial Profiles of Methanol Removal .............................................................................................................


56

4.1

4.1.2

Biomass Accumulation in the Packed Be& ............................................................. 57

Biomass Quantificationas Detached Biolayer.............................................................................................. 5 8


Biomass Quantification as Chcmical Oxygen Demand (COD).Total Solids (TS). and Total Volatile
Solids (TVS) ...........................................................................................................................................
6
0
Patterns of Bioinass Accumulation ............................................................................................................... 63

Pressure Drop (A?') .................................................................................................64


Total Pressure Drop .......................................................................................................................................
64
Pressure Drop Axial Profiles ........................................................................................................................
66
Pressure Drop vs.Airtiow Rate.................................................................................................................. 6 7

Bcd Compaction ......................................................................................................................................... 69

Bed Moisture Content and Interparticle Porosity ...................................................70


SWITCHOVER
FROM METHANOL
TO a-PINENE ..................................................................
73
Switchover Period ....................................................................................................73
Overall Performance, Pressure Drop, and Biomass Growth.................................. 7.1
TRACER
EXPERIMENTS
.....................................................................................................
78

EVALUATION
OF BIOMASS
QUANTIFICATION
TECHNIQUES...............................................
79
PRESSUREDROPAND CUMULATIVE
METHANOL
REMOVAL .............................................
82
PRESSUREDROPAND BIOMASS
ACCUMULATION
DURMG THE BIOFILTRATION
OF METHANOL
.................................................................................................................84
Correlation between Pressure Drop and Biomass Accumulated in the beds .......... 84
Biomuss Distribution Effects on Pressure Drop .................................................... 88
COMPARISON
OF PERFORMANCE
BETWEEN PACKING
MEDIA...........................................89
Pollutant Removal ...................................................................................................89
Pressure Drop .......................................................................................................... 92
Biomass Accumulation.............................................................................................93
.4lRFLOW CHARACTERISTICS.....~~......................................................................................
94

Correlation between Pressure Drop and Airfow raie ............................................ 94


MODELLING
PRESSUREDROPvs. BIOMASS
ACCUMULATION
..........................................95
Mode1 Development ................................................................................................96
Estimation of Biofilm Thickness...................................................................................................................96
Estimation of Biornass-affectcd Porosity................................................................................................... 9 7
Estimation of Pressure Drap..........................................................................................................................98

Cornparison beween merimental and Estimated Pressure Drop ........................ 99


Cornparison among Models ................................................................................... 102

NOMENCLATURE .........b~........ama.a.m.a.m.am.ma..m........m..........m......m..a..............m.........
79

REFERENCES ......b................a..m.m..aam.mmmam.m....mm......m.....mam.a...a..m......mm.......m...
111

APPENDICES
A. Calculation of Nutrient Requirements for Both Biofilters

B. Calibration Data for the Varian 3600CX Gas Chromatograph

C. Biomass Quantification Profiles for the Nova Inert and Wood Chop Biofilter
as COD, TS, and VS during the biofiltration of methanol
D. Cornparison between Biomass Measurernents as COD/TS and Detached
Biomass in the Wood Chip Biofilter

E. Transient Conditions after Bed Sampling and Remixing, before Switching


over to a-pinene
F. Transient Conditions aller Switching over fiom Methanol to a-pinene
G. Pressure Drop Profiles during the Biofiltration of a-pinene
H. Biomass Quantification during the Biofiltration of a-pinene
1. Estimation of the Yield Coefficient for the Nova Inert and Wood Chip

Biofilter
J. Calculation of Reynolds Numbers (NRE)b
for Both Biofilters

K. Pulse Tracer Tests


L. Calculation of Adjustable Panuneters N (number of tanks in series) and PeL
(Peclet number)

M.Modelling Pressure Drop

as a Function of Biomass Accumulation-

Calculations

vii

LIST OF FIGURES
Page
Figure 2.1
Figure 3.1
Figure 3.2
Figure 3.3
Figure 3.4
Figure 3.5
Figure 4.1
Figure 4.2
Figure 4.3

Figure 4.4

Figure 4.5

Figure 4.6
Figure 4.7

Figure 4.8

Figure 4.9
Figure 4.10

Relation of factors affecting biofilrn growth and its accumulation (Van


Loosdrecht et al., 1995)
Schematic representation of the experimental set-up for one biofilter
Photograph of the actual experimental set-up. From lefi to right: control
panel, valve control panel and manometer for AP measurement,
humidification system, biofihers, and GC
Photograph of section 1 of the wood chip biofilter on day 1 1 that shows
the axial ports for pressure drop measurement on the right hand side
Photograph of section 1 of the Nova Inert biofilter on day 1 1 that shows
the axial ports for pressure drop measurement on the right hand side
Sampling points for composite samples at the top, middle, and bottom of
each section of the biofilters
Idet and outlet concentrations of the Nova Inert and wood chip biofilter
during the biofiltration of methanol at 40C bed temperature
Nova Inert and wood chip biofilters performance during the biofiltration
of methanol at 40 O C bed temperature
Typical methanol axial concentrations dong the Nova Inert biofilter
before sampling and remixing the bed during the biofiltration of
methanol. A dashed line indicates the approximate location of each
section
Typical methanol axial concentrations along the wood chip biofilter
before sampling and remixing the bed during the biofiltration of
methanol. A dashed line indicates the approximate location of each
section
Biomass concentration profiles as detached biolayer along the Nova Inert
biofilter in 4 dinerent days during the biofiltration of methanol. A
dashed line indicates the approximate location of each section.
Biomass concentration profiles as detached biolayer dong the wood chip
biofilter in 4 different days during the biofiltration of methanol
Comparison between biomass measurements as mg CODIg dry
substratum and as g biolayerlg dry substratum in the Nova Inert biofilter
during the biofiltration of methanol
Comparison between biomass measurements as mg TS/g dry substratum
and as g biolayedg dry substratum in the Nova Inert biofilter d u ~ the
g
biofikration of methanol
Photographs of top levels of the four sections of the Nova Inert biofilter
at the end of the second undisturbed-bed period of operation (day 84)
Photographs of top levels of the four sections of the wood chip biofilter
at the end of the second undisturbed-bed period of operation (day 74)

viii

Figure 4.1 1

Figure 4.12
Figure 4.13
Figure 4.14
Figure 4.15
Figure 4.16
Figure 4.17
Figure 4.18
Figure 4.19
Figure 4.20
Figure 4.21
Figure 4.22
Figure 4.23
Figure 4.24
Figure 5.1
Figure 5.2

Total pressure &op along the Nova Inert and the wood chip biofilten
during the biofiltration of methanol. Bed remixing points are indicated
by continuous lines for the Nova Inert bed and by dotted lines for the
wood chip bed.
Cumulative pressure drop profiles dong the Nova Inert biofilter at the
beginning and at the end of each undisturbed-bed period of operation. A
dashed line indicates the approximate location of each section
Cumulative pressure &op profiles along the wood chip biofilter at the
beginning and at the end of each undisnirbed-bed period of operation. A
dashed line indicates the approximate location of each section
Total pressure drop in the Nova Inert biofilter, as a function of aifflow
rate, measured on different days during the third undisturbed-bed period
of operation
Total pressure drop in the wood chip biofilter, as a function of airflow
rate, measured on different days during the third undisturbed-bed period
of operation
Percentage bed compaction measured as percentage of initial bed height
in the wood chip biofilter during the four undisturbed-bed operation
periods. The duration of each period appears in parenthesis on the x axis
Moisture content in each section of the Nova Inert biofilter at the end of
each undisturbed-bed period of operation.
Moisture content in each section of the wood chip biofilter at the end of
each undisturbed-bed period of operation.
Interparticle porosity in each section of the Nova Inert biofilter at the end
of each undisturbed-bed period of operation
Interparticle porosity in each section of the wood chip biofilter at the end
of each undisturbed-bed penod of operation.
Performance of the Nova Inert and wood chip biofilters removing apinene, after 150 days of operation with methanol
Total pressure drop in the Nova Inert and wood chip biofilters during the
biofiltration of a-pinene
Profiles of biomass quantification as detached biolayer along the Nova
Inert biofilter and the wood chip biofilter at 2 bed sampling days during
the biofiltration of a-pinene
Cumulative pressure drop profiles along the wood chip biofilter at the
beglluiing and at the end of each undisturbed-bed operation period during
the biofiltration of a-pinene
Comparison between biomass measurements as mg CODIg dry
substratum and as mg TS (TVS)/g dry substratum in the Nova Inert
biofilter during the biofiltration of methanol
Comparison between biomass measurements as mg CODIg dry
substratum and as mg TS (TVS)/g dry substratum in the wood chip
biofilter during the biofiltration of methanol

Figure 5.3
Figure 5.4
Figure 5.5
Figure 5.6
Figure 5.7
Figure 5.8
Figure 5.9
Figure 5.10

Figure 5.1 1
Figure 5.12

Pressure drop along each section as a fnction of the cumulative amount


of methanol removed in each section of the Nova Inert biofilter,
measured at the end of each undisturbed-bed operation penod
Pressure drop along each section as a function of the cumulative amount
of methanol removed in each section of the wood chip biofilter,
measured at the end of each undisturbed-bed operation period
Pressure drop along each section as a function of the average arnount of
biomass accumulated in each section of the Nova Inert biofilter,
measured at the end of each undisturbed-bed operation period
Pressure drop along each section as a huiction of the average amount of
biomass accumulated in each section of the wood chip biofilter,
measured at the end of each undisturbed-bed operation period
Pressure drop along sections 3 and 4 as a function of percent compaction
in each section of the wood chip biofilter, measured at the end of each
undisturbed-bed period
Comparison of the performance of the Nova Inert and wood chip
biofilters in terms of removai capacity in each section of the biofilten
Experimental and estimated by the proposed model pressure drop values
for the Nova Inert biofilter during the biofiltration of methanol
Comparison between wet porosity measured fiom well mixed bed
samples and the bed porosity in the Nova Inert biofilter as predicted by
the M-E equation fiom AP experimental data, and by the biomass growth
model on spheres in contact with each other
Comparison between values of pressure drop estimated by the explained
models and the experimental values fiom the Nova Inert biofilter
Comparison between estimated pressure drop from averaged biomass
concentrations and experimental pressure drop data before and d e r
remixing the Nova Inert bed

LIST OF TABLES
Table 2.1
Table 2.2
Table 3.1
Table 3.2
Table 3.3

Some odorous, volatile organic, and toxic compounds that are emitted by
the pulp, paper, and wood industries, and that have been treated at
different scales by biofiltration
Some indirect techniques for biofilm and biomass quantification
(Adapted fiom Characklis (1990) and Tate III (1 995))
Size distributions in percentage volume for wood chips and Nova Inen
pellets used to pack the biofilters
Composition and some specifications of the utr ri cote^ slow-release
fertiliser pellets used as a source of macronutrients in the biofilters
Composition of micronutrient solution added to the Nova Inert packing
material

6
24

43
44

45

Introduction

1 INTRODUCTION
1.1 The Biofiltration Process

Biofilters, along with biotrickling filters and bioscnibbers, are biological treatment units.
which are increasingly being used as air pollution control techologies. A biofilter is a multiphase, packed, biological reactor whose operation involves venting a polluted air stream through
a packed bed of mostly natural matenals, which under optimal conditions of humidity,
temperature, a .pH, permits a microbial consortium to develop and degrade the pollutant.
Biofiltration is technically and economically efficient for the treatment of moist, dilute
air streams contaminated with odorous, toxic, and volatile organic compounds. Therefore, it has
a potential application to industries such as the pulp and paper industry, where these types of
strearns can be found.

Research has been conducted on several topics related to the

performance of biofilters: different support materiais, optimal particle size, bed inoculation, and
microbial community development. This research has allowed the implementation of bio filters
to solve curent pollution control problems; however, the process c m be optimised by better
understanding the phenomena taking place in the biofilter bed. These phenomena include the
development of a degrading microbiai cornmunity and its effects on the airflow characteristics
through the biofilters, specifically, on the increase in pressure &op or air resistance through the
bed.

1.2 Statement of Research


1.2.1 Problem and Hypothesis
A better understanding of the basic mechanisms that affect the operation of a biofilter

will help enhance performance, for example by reducing energy requirements, and operating
and maintenance costs. In particular, it is necessary to investigate the general problem of how

the development and accumulation of biomass on the support material influences pressure losses

in a biofilter that treats odorous compounds aad/or volatile organic compounds. The following
general hypothesis can be proposed to answer this question: biomass growth in a biofilter
increases pressure drop by reducing the interparticle void space in the bed, which leads to partial

Effects of Biomuss Growth on Pressure Drop in Biofilrers

Introduction
clogging and channelling, and this increased pressure drop can be descnbed as a function of
biomass growth and substrate consumption.
1.2.2 Objectives

The objectives of this research project were to study and quanti@ the influence of
different levels of biomass accumulation on pressure drop in biofilters, and to model pressure
losses caused by biomass growth in biofilters using semi-empirical models such as the
Modified-Ergun equation and specific permeability models for flow through porous media.
These objectives were achieved in two stages. First, two bench-scale biofilters were
operated for removing methanol and a-pinene. One was packed with wood chips and the other
with an inert material. Then, a phenomenological model was proposed for predicting pressure

drop and interparticle porosity changes due to biomass accumulation in a biofilter packed with
inert spherical pellets.
1.2.3 Significance

In the process of biofiltration, as the biodegradable pollutants are removed from the
contaminated air Stream, a microbial cornmunity accumulates on the filter material. The
synthesis of new cells by the mineralisation of organic matter leads over time to biomass growth
and to a change in the filter bed charactenstics. Furthemore, reduction of the interparticle void
space in the bed by biomass development and clogging, together with bed compaction, causes
an increase in flow resistance.
The pressure loss in a biofilter depends on different operational parameters and
represents a long-terni disadvantage for full-scale biofilters. As pressure drop increases with
biomass accumulation and with the gradua1 compaction of the natural filter bed, more power
will be needed, and the eventual requirernents of mixing or replacing the filter material will
cause higher operational and maintenance costs. The effects of biomass growth on pressure
&op in biofilters have already been observed and reported in several experimental studies.
However, no specific studies have addressed this problem in order to investigate, and to model

the fiuidarnental xnechanisms that determine the pressure drop in biofilters in relation to biomass
accumulation. This knowledge will help to identify sources of operational problems and

measures for improving the biofiltration process.


Effects of Biomass Growth on Pressure Drop in Biofiifers

Introduction

1.3 Thesis Outline

nie body of this thesis is divided into the following parts. After this brief introduction, a
literature review is presented in the following section (section 2). The theoretical framework of
this study covers a description of the biofiltration process and its applicability to the pulp and
paper industry.

The parameters afTecting pressure drop in a biofilter, such as the bed

characteristics and biomass accumulation, are also reviewed. Parts of section 2 are dedicated to
the process of biomass development in filter packed beds, its distribution and quantification, and
its effects on pressure drop. Modelling of the effects of biomass growth on pressure drop are
also studied for the cases of groundwater flow in porous media and bioreactors.
Sections 3 presents the experimental approach, including the biofiltration system set-up,
procedures, and analytical techniques involved in this project. Section 4 is dedicated to the
results, which cover biofilter performance, observations of biomass accumulation and pressure
drop.

The discussion of results (section 5) is focused on the evaluation of biomass

quantification techniques, on the analysis of pressure drop and biomass accumulation and
methanol removed in both biofilters, on the airfiow charactenstics and traces tests, and finally
on the modelling of pressure drop as a fiction of biomass accumulation. After an explanation

of the hdamental equations of the model, the predicted and experimental values of pressure
drop are compared. Final conclusions and recommendations are given in section 6. Several
appendices present details regarding raw data, specific results, and calculations.

Ef/ects of Biomass G m t h on Pressure Drop in Biofilters

Theoretical Background and Literature Review

2 THEORETICAL BACKGROUND A N D LITERATURE REVIEW


2.1 Biofiltntion and Its Application to the Treatment of Pulp, Paper, and Wood

Air Waste Emissions


2.1.1 The Process of Biofiltration

Biofiltration is a biological air purification technology that has been applied to treat a
variety of odorous, volatile, and toxic compounds. In a typical biofilter, the contarninated air
Stream is passed through a packed bed of porous materials, mostly organic, that serve as a

support for a microbial community that, under optimised conditions, degrades the incoming
pollutants into carbon dioxide, minera1 salts, water, and microbial biomass.
Biofiltration, initially used for the treatment of odorous compounds, has been extended
to the treatment of volatile organic compounds (VOC's) and toxic compounds (Ottengraf et al.,
1986; Leson and Winer, 1991). In the past decade, extensive reviews and snidies about the

development and technical aspects of biofiltration have been published (Leson and Winer, 1991;
Togna and Singh, 1994; Swanson and Loehr, 1997; Wani et al., 1997).
Motivated by new legislation, such as the 1990 Clean Air Act amendments in the U. S.?
significant fundamental and applied research on biofiltration has taken place. In addition,
biofiltration has been considered economically advantageous for the treatment of large flow air
waste streams, typically from 1000 to 50,000 m3/h, containing low concentrations of odorous,
toxic and VOC biodegradable pollutants, in a concentration typically less than 1000 ppmv as
methane (Leson and Winer, 199 1; Devinny et ai., 1999). Biofiltration is possible at relatively
low gas residence times, providing economic cornpetitiveness, especially at gas residence times
less than 45 seconds (Stadefer and Willingham, 1998). Unlike conventional technologies, such

as thermal and catalytic incineration or carbon adsorption, biofiltration allows effective


pollution control at a relatively low capital and operating costs, and without generation of
secondary streams that may need a subsequent treatment.

Effectsof Bionuss Growth on Pressure Drop in Biofilers

Theoretical Backaround and Literuture Review


2.1.2 Application to the Pulp, Paper, and W

d Industries

Pulp and paper mills contribute to industrial air pollution with the discharge of odorous,
volatile organic, and toxic air compounds. For example, reduced sulphur compounds are
produced in kraft pulping and the recovery process (Jones, 1973), chloroform and methanol are
generated in the bleaching process, and formaldehyde can be emitted fiom papermaking and
fiom certain wood processing operations. Alcohols, terpenes, and other volatile organic
compounds (VOC's) are also released nom bleaching and wood processing units (Jain, 1996).
Even though process changes and the use of conventional air pollution control technologies
have reduced air pollutant emissions fiom pulp and paper mills, significant arnounts of odorous,
volatile organic, and toxic compounds are still released, specificaily in already treated off-gas
streams and fugitive emissions (Leson et al., 1991).
This situation together with stticter regulation such as the Maximum Achievable Control
Technology (MACT) standards, derived fiom the 1990 Clean Air Amendments, have
encouraged emission reduction and control in the pulp and paper industry. In fact, in the case of
the U.S.,the 1997 Phase 1 of the U.S. Environmental Protection Agency's (EPA) Cluster Rule,
establishes national emission standards for hazardous air pollutants (HAP's) at al1 chernical
pulping and bleaching facilities in the country (Pinkerton, 1988; Vice and Carroll, 1998).
Among these HAP's are certain VOC's, such as methanol, that have been identified by the EPA
as compounds of significance for chernical pulp mills (Dence and Reeve, 1996) because of their
carcinogenic characteristics and their role as primary precursors of ground-level ozone, which is
related to cardiorespiratory health affections.
Compound types that are typically degraded by biofiltration include alcohols, esters,
aldehydes, ketones and common monocyclic aromatics, several nitrogen-, sulphur-containing
organic compounds, and to a lesser extent higher chlorinated organic compounds (Leson and
Winer, 1991; Devinny et al., 1999). Coincidentally, most of these compounds are present in
waste air emissions fiom the pulp, paper, and wood industry. Table 2.1 shows some compounds
that are observed in pulp and paper mills off-gas streams, and that have been successflly
treated by biofiltration.

The treatability of these pollutants by biofiltration but also its

advantages for treating large flow low concentrated air streams render biofiltration potentially
applicable to this type of industrial waste emissions.
Effects of Biomass Growth on Pressure Drop in Biofilters

Theoretical Background and Literaiure Review

Some odorous, volatile organic, and toxic compounds that are emitted by the
Table 2.1
pulp, paper, and wood industries, and that have been treated at different scales by biofiltration

Reduced sulphur
compounds:
hydrogen sulphide (H2S),
methyl mercaptan
CH3SH),
dimethyl sulphide
(CH3SCH3)

Xy lene
Mixtures of benzene,
toiuene, and xvlene
Mixture of ethylacetate,
butylacetate, butanol, and
toluene

Methyl ethyl
ketone(MEK) and methyl
isobuthyl ketone (MIBK)
Mixture of VOCYs:
benzene, toluene,
dichioromethane,
trichioroethene,
tetrachloroethene, and
trichioromethane

Reported biofilter
application
General source
Various stages in the Lab experiments,
Kraft,sulphite
wastewater plants,
pulping and the
solid waste
recovery process
processing plants,
rendering plants,
chernical
manufacturing and
composting
operations

Reference
Cho et al., 1991;
Shoda, 1991 (Van
Langenhove et al.,
1986; Allen and

Yang, 1991; Williams


and Miller, 1992b);
Sabo, 1989; Paul et
al., 1989 (Leson, et
al., 1991 ; Ergas et al.,
1995)
Released in pulping Laboratory studies (Mohseni, 1998);Van
and bleaching
Lith et al.,
systems
1990(Williarns and
Miller, 1992b;
Shareefdeen rr al..
1993)
Generated in smaller Laboratory study
(Bardtke et al., 198 7)
quantities in pulp
Laboratory studies (Corsi and Seed,
bleaching and wood
1995)
processing
Laboratory study
Ottengraf and Van
den Oever, 1983
(Corsi and Seed,
1995)
(Hodge and Devinny,
1995)
Laboratory study
(Deshusses et al.,
1995)

Laboratory study

(Ergas et al., 1995)

Laboratory study

[Apel et al., 1995);


[Mohseni and Allen,
1997)

Efects of Biomass Growth on Pressure Drop in Biojilfers

Theoretical Backaround and Literature Review

Bench-scale
laboratory
experiments
Laboratory study

(Mohseni, 1998)

Wood processing
plants

~ a b o r a t 0 6study

(Coleman and
Dombroski, 1996)

Wood products
industry

Full-scale biofilter

Press vent emission


fiom an oriented
strand board
manufactwing plant
Particle press vent

Full scale
biofiltration
system

Methmol and a-pinene


Mixture of MEK,
methylmercaptan,
dimethyl sulphide,
hydrogen sulphide,
methanol, acetone,
turpentine, and ethanol
(representative of a
digester blowoff)
Mixture of acetic acid,
formaldehyde, and
methanol
Mixture of VOC's
(benzene, toluene,
alcohols, aldehydes, and
organic acids) and other
air toxics
Odorous compounds

Mixture of formaldehyde,
a-and h i n e n e
Table 2.1 continued

Wood fiber dryers


and Kraft pulphg

(Sinitsyn et al., 1991)

1994 (Wani et al.,


1997)

Full scale biofilter

(Devinny et al., 1999)

2 3 Parameters Affecting Pressure Drop in Biofilters


Energy consumption is a substantial part of the operating costs of a biofiltration system,
and the majonty of the electricity requirements are devoted to the operation of the air blower.
In a packed bed, the pressure drop is caused by viscous and kinetic energy losses, and it depends
on the fluid characteristics and on the packing material properties (Ergun, 1952). Thus, the
pressure drop across a biofilter can be descnbed as a fnction of the aifflow characteristics, such

as flow rate, air density and viscosity, and as a function of the properties of the packing
medium, namely, particle size and fractional void volume or porosity .

The pressure drop through a biofilter bed ranges typically fkom 20 to 98 Pa (0.08 to 0.4
in of water), and can even go up to 980 Pa (4 in of water). Typical superficial gas velocities in a
biofilter may vary fiom 5 to 500 m3Im2 h (Devinny et al., 1999). Gas velocities of up to 300
Eflects of Biomuss Growih on Pressure Drop in Biojilrers

Theoretical Background a ~ Literature


d
Review

m3/m2 h, or even as high as 500 m3lm2 h in an optimised bed, are usually feasible without
resulting in excessively high pressure drops

(C

1O00 Pa) (Leson and Winer, 1991). Systems

with adequate moisture control and a porous biofilter medium containing bulking agents will

achieve pressure losses of typically less than 900 to 1700 Palm (Leson and Smith, 1997).
The main variables that have been identified to influence pressure drop in a biofilter are:
1. Superficial gas velocity (m3~ff-~adrn~filter
h or m3/m2 h )

2. Characteristics of the packing material

a) Particle size distribution and bed depth

b) Nature and composition of the packing material

c) Aging of the packing material


3. Bed moisture content

4. Biofilm development andor biomass growth

Even though each of these factors may affect pressure drop in a biofilter in a separate
way, it is the combined effects of the flow characteristics with the properties of the filter bed,

which determines pressure losses in a biofilter. In the following sections these factors are
discussed in more detail.
2.2.1 Superficial Gas Velocity

In a biofilter, as in any packed bed, the higher the superficial gas velocity, the higher the
pressure drop dong the bed. However, the bed depth and the way the filter medium is packed,
which is described by its bulk density and porosity, determine the pressure drop across the bed
at a given gas velocity.

The relationship between pressure &op and superficial air velocity or airfiow rate has
been reported by several authors (Leson and Winer, 1991;Hodge et al., 1992), mostly in media
characterisation studies applicable to biofiltration. Sabo et al. (1993) conducted experiments
with different filter materials where pressure &op was determined at various gas flow rates and

at diflerent bed moisture contents. For the tested materials (bark compost/porous burned clay,
wood chips, and coconut fibres), the pressure drop first increased in direct proportion to the

rising air velocity, but it deviated from linearity at higher air velocities. Pressure drops as high

as 3000 Pa were registered in the 2.5 m bark compost~porousburned clay column, and as high
Emts of Bioniass Growth on Pressure D ~ k
GBiofilers

Theoretical Background and Literature Revie w

as 300 Pa in the 2.5 m coconut fibre and wood chip columns, using superficial air velocities up

to 550m3lm2 h.
Analogous observations were obtained by Van Langenhove et al. (1986) working with
wood bark and fibre peat filter media, and by Allen and Yang (1991) using different kinds of
compost. Although al1 these studies report an increase in pressure drop with increasing gas
velocity, how much the pressure &op increases and at which air velocity there is a transition
fiom a linear to a parabolic behaviour seem to depend on the type of material used.
Semiparabolic curves of pressure drop vs. superficial air velocity, with a linear
behaviour at low air velocities, have been widely reported for natural porous packing materials
(Madamba et ai., 1994). These types of curves can be satisfactorily modelled by a Modified
Ergun (M-E) equation which takes into account the dependence on porosity of the viscous and
kinetic energy losses (Macdonald et al., 1979).

Macdonald et al. (1979) proposed a

modification to the original quasiempirical Ergun equation (Eq. 1), which is applicable to any
type of flow, in order to account for particle roughness and a better porosity function in
accordance to flow through porous media. From the Ergun equation (Ergun, 1952):

the constants 150 and 1.75,and E' were substituted by A, B, and E

~ *respectively,
~ ,

as presented

in the M-Eequation:

AP

(I-EY~u,

=A----.L gc

3.6

(I-E)GU~~,
BE>.^op.
DP
(Eq*2)
*

Here AP is the pressure drop across the bed, L is bed height, g, is the gravitational constant, s is
the bed fiactional void volume, p is the air absolute viscosity, U, is the superficial air velocity,

G is the air mass flow rate (G=pU),

4 is the effective diarneter of particles, and A and B are

constants. Panuneter A is considered to be independent of particle roughness and equal to 180,


but B is dependent on surface roughness varying fiom 1.8 for smooth particles to 4.0 for the
roughest particles. The first term in the M-E equation takes into account viscous energy losses,
proportional to

( ~ - E ) ~ / E " ~and
,

the second term accounts for kinetic energy losses, proportional

to (1-E)IE~-~.

- --

Effects of Biomess Growth on Pressure Drop in Biofdters

Theoretical Background and Literature Review

10

Without considering the media aging phenomenon, the use of the Ergun equation has been
suggested as a rough estimate of pressure drop as a function of superficial air velocity in
biofilters (Ottengraf et al., 1986). However, the M-Eequation might be useful for the same
scenario in biofilters since it takes into account the properties of packed beds of porous
materials, including natural media. On the other hand, the prediction of headlosses through a
biofilter still represents a challenge due to the high variability of the physical characteristics of
packing media, and their property changes over operation time. The M-E equation has been
used to mode1 pressure drop data for compost-wood chip packing mixtures in a biofilter without
biomass growth (MacFarlane, 1998).
2.2.2 Bed Characteristics

The bed characteristics of a biofilter determine its eficiency of operation. The filter
medium has been considered as the heart of the system since it dictates removal eficiency, air
distribution, moisture content, and pressure &op. Among the requirements that the packing
material should meet are: optimal environmental conditions for microbial degradation (supply of
nutrients, pH, moisture), large specific surface area that maximises microbial attachent and
growth area, adequate moisture content, structurai resistance to avoid compaction (void space
reduction), low pressure drop and high porosity, low cost, availability, and ease of disposal.
Most of these characteristics are intimately related to the flow resistance through the bed as
outlined below.
2.2.2.1 Particle Size Distribution and Bed Depth

It is well established that in a biofilter, as in any packed colurnn, very small particles
cause high pressure drops since they tend to form a compact bed with high bulk density, and
lower specific penneability due to packing arrangements with smaller channel diarneter. For
biofilters, it is recornmended that 60% (by weight) of fiter particles have a diameter greater

than 4 mm (Corsi and Seed, 1995). Particles with an effective diarneter equal to 4 mm and with

a uniform shape (cylindrical carbon pellets, for example) may still induce greater pressure losses
than particles with a similar diameter but with irregular shapes (e.g. perlite) under the sarne
conditions (Mohseni et al., 1998). Smaller particles of the filter medium (diameters less than 12 mm) have been observed to cause a signifcant increase in pressure drop, compared to bigger
Effects of Bbmass Growth on Pressure Drop in Biofiiters

Theorefical Background and Literature Review

11

particles at the same gas velocities (Van Langenhove et al., 1986; Allen and Yang, 1991).
Media with a high proportion of particles less than 3 mm in diameter may cause pressure drops
in excess of 1800 to 3300 Pdm, even at residence times of one minute or more (Leson and
Smith, 1997). This increased headloss may be due partially to the presence of small particles
such as sand or decomposed and mineralised organic matenal generated as a result of fracture
and medium aging. In general, it is recommended to avoid the smallest particles by sieving the

medium, and to use a particle size that provides a reasonable surface area and an acceptable
flow resistance.
Mixing small particles (diameter = 4-6 mm) with large particles (diameter > 10 mm) of
media have been reported as a way of reducing pressure losses through biofilters (Ottengraf et
al., 1986). Pressure drop also increases approximately linearly with bed packing height.
2.2.2.2 Nature and Composition of the Filter Medium

The necessary requirements of a packing medium may be met by mixing different


materials. Natural materials, such as compost, bark, peat, and wood chips, have been used as
packing materials since they satisfy several requirements of a biofilter medium; however, they
are subject to aging phenomena, which have been related to an increase in pressure drop. To
lessen compaction, increase durability, prevent channelling, increase surface area and void
space, and thus reduce pressure drop, nahual media cm be mixed with inert materials, like
porous clay, activated carbon, polystyrene spheres, coarse sand, and perlite. A proper selection
of the filter material will help to reduce pressure drop through the bed.
Shareefdeen et al. (1993) observed that the addition of porous materials, such as
vermiculite and perlite, to peat decreased the pressure drop in a peat packed colurnn. The
mixture ratio chosen was 40% peat and 60% perlite (vlv). The peat-perlite mixture performed
better than either ingredient alone. Apparently, peat, a good microbial support, has pore spaces
too small for even aeration, leading to channelling and poor contact with the gas phase. The
addition of perlite assured that more peat particles had good contact with the gas phase, causing
the superior performance of the mixture. Corsi and Seed (1995) used the same ratio (60:40)for
mixtures of perlite and compost, while Mohseni et al. (1998) used a 75/25 ratio for mixtures of
wuod chips/compost and perlite. Mohseni and Allen (1997) reported removal efficiencies

Effects of Biomass Growth on Pressure Drop in ~ i o f i l t = i

12

Theorerical Backaround and Literature Review

higher than 95 % for most of the operation penod (130 days) and suggested that the high
removal efficiencies attained could have been a partial result of the high surface area per unit
volume of the bed obtained by the addition of the inert rnaterials. The pressure drop through the
perlite biofilter was reported to be initially very low (less than 45.5 Pa/m during the first 50 days
of operation) and later not higher than 135 Pdm, with a constant empty bed retention time

(EBRT)of 45 sec. Higher pressure drops (270-365 Pa/m) were reported at higher gas flow rates
(lower retention time EBRT = 30 sec). The addition of porous materials with high intemal
porosity and with hydrophilic characteristics may also function as a buffer for excess moisture
in the bed; thus, decreasing headlosses (Ottengraf et al., 1986).
Van Langenhove et al. (1986) reported that wood bark gave less pressure drop than fibre
peat or household compost. In general, compost and soi1 cause the highest pressure drops (8001500 Palm;

0.3-0.4) arnong the natural materials used in biofilters (Bardtke et al., 1987).

Although peat has a low pressure drop (100 - 300 Pdm;

= 0.7-0.85), its use has decreased

because of difficulty controlling moisture in peat beds and low long-term performance ( D e v i ~ y
et al., 1999). Several types of compost present useful properties for biofilters, e.g. large

microbiai comunity and nutrient supply, and new engineered media such as pelletised compost
may render the use of compost with lower pressure drops (Zich et al., 1996; Devinny et aL,
1999). Wood chips and bark have been used as bulking agents and have led to lower pressure

drops in biofilters, yet there are not many studies using these media alone in biofilters (Devimy
et al., 1999).

In general, the improvement of natural media characteristics by the addition of


amendment materials is a common practice, which is explained by a reduction of the bed bulk
density, an increase in porosity, and a decrease in pressure drop and compaction.
2.2.2.3 Moisture Content

The moisture content of a biofilter bed is a critical parameter since it affects the
metabolism of the rnicroorganisms, the properties of the packing material, and therefore, the
pressure drop across the bed and the general efficiency of the process. Depending on the
packing material used, the optimal bed moisture content rnay Vary from 30 to 60% on a weight
basis (Williams and Miller, l992a). Several authors have reported increases in pressure losses
Efjects of Biomass Growth on Pressure Drop in Biofiters

Theoretical Background and Literature Review

13

through biofilters with increasing bed moisture content. Sabo et al. (1993) concluded that bed
moisture content was the most important parameter, compared to superficial air velocity,
determining the transition region fiom laminar to turbulent flow in biofilters packed with bark
compost/ porous burned clay (50150 vol%), coconut fibres, or wood chips. In addition, they also
reported the influence of the packing matenal structure on the moishue content, indicating that
increasing the moisture content in coarse fibre materials did not increase pressure drop as much

as in granular stnictured materials. The strong relation between moisture content and pressure
drop dong a biofilter bed suggests that pressure drop may be a useful indicator of bed moisture
content in biofilters (Van Langenhove et al., 1986).
The effects of insufficient or excessive bed moisture content on the flow characteristics
in biofilters have been widely cited in the literature (Wani et al., 1997; Wright et al., 1997). An
excess of moisture causes the gas filled porosity of the bed to decrease since water occupies the
void pores within the bed; therefore, increasing headlosses in the system. High bed moisture
promotes clogging, nutrient leaching, accelerated compaction, and also the creation of anaerobic
zones that may cause odour problems and oxygen transport resistance resulting in a decrease of
the biodegradation efficiency. On the other hand, a low bed moisture content will result in
drying of the filter material, which could lead to cracking and channelling within the bed. This
would provoke higher pressure drop, and a decrease in retention time and, hence, in the removal
efficiency. Bohn (1976) (Corsi and Seed, 1995) reported that bed drying and cracking can occur
at high gas flow rates. Similarly, Barshter et al. (1993) (Corsi and Seed, 1995) observed bed
drying due to locdly high airfow rates in a full-scale biofilter. These effects were most
fiequent near the corners and dong the walls of the vesse1 containing the bed material.
Moishue to the biofilter bed can be supplied by prehumidified air (no less than 99% RH)
orland by sprinkling water at the top of the filter. Unnecessarily high pressure drop has been
recorded when sprinkling water on the top of the filter, and it seems preferable to condition the
relative humidity of air entering the filter in order to moisten the filter bed (Van Langenhove et
al., 1986), and even control its temperature. Also, water spraying does not allow for uniform
wetting of the packing media (Utkin et al., 1989).

Effects ufBiomass Growth on Pressure Drop in BiofiIters

Theoretical Background and Literature Review

14

2.2.3 Microbial Mass Growth

Over time in most biufilters, as biomass begins to grow on the packing material, the
pressure drop begins to increase. Govind et al. (1993) observed that in an activated carbon
biofilter there is a point in time at which the pressure drop begins to increase and goes steadily
up. While sorne authon do not mention significant pressure losses during biofiltration studies
and do not report any influence of biomass on the head losses across the biofilters, some others
do relate increases in pressure drop to biofilm growth and clogging by biomass (Utkin, et al.,
1989; Hodge, et al., 1992). The increase in pressure &op by biomass development in biofilters
cm be explained by

a decrease in the bed void space, and the microbial degradation of the

support matrix, as in the case of natural media.


Liu et al. (1994), working with granular activated carbon (GAC) columns for toluene
removal, noticed increases in filter pressure drop due to biomass accumulation and airflow
channelling. Biomass occupied the interparticle spaces within the bed, and its partial removal
was necessary in order to maintain a low pressure drop. The pressure drop in the columns
packed with GAC (2.54 cm in diameter and 55 cm in height) varied fiom O to 5 kPa with
airflow rates ranging from 0.5 to 4 Vmin and inlet toluene concentrations of 10 to 20 ppm.
Increased pressure drops up to 6.9 kPa were reported to result fiom biomass build-up in the inlet
section and fiom channelling. Any specific pattern of pressure drop increase and of biomass
accumulation was not reported. A subsequent study using a pilot-scale unit with a larger
diameter and with automatic biomass removal exhibited similar or lower pressure drops (Liu et
al., 1994).

Although the effect of biomass accumulation on pressure drop in biofilters has been
reported, there has not been detailed quantitative research of this phenornenon. Thus, this
constitutes one of the objectives of this study.

2.3 Biomass Accumulation and Pressure Drop in Filter Packed Beds


2.3.1 Biomass Development in Bio%ters

Biomass in a biofilter medium grows as a biofilm, which consists of a mixture of


microbial ceiis, extracellular polyrners, lysis and hydrolysis products, attached organic matter,
-

Effects of Biomass Growth on Pressure Drop in Biofilers

15

Theoretical Background and Literature Review

and some inorganic compounds, al1 of them bound tg an inanimate solid surface, termed

substratum. The predominant microorganisms in biofilms of biofilters treating VOC's are


heterotrophic, which include many types of bacteria and fungi. In a compost biofilter, Helmer
(1972) (Ottengraf et a l , 1986) detected white material with growing fungi, more frequently in

the lower layers of the filter bed. Other organisms fiequently found in biofilters include
protozoa and nematodes (de Castro et al., 1997). While bacteria have higher rates of metabolic
activity and usuaily predominate in aqueous media, fungi are less affected by low moisture
environrnents and can grow on dry media using moisture fiom the air.
The term biomass growth or biomass accumulation involves a collection of metabolic
processes such as substrate conversion, cellular growth, replication, endogenous decay, and
extracellular polymenc substance (EPS) production.

The terni EPS not only includes

microbially produced bound polymers but also lysis and hydrolysis products and adsorbed or
attached matter (Nielsen et al., 1997).
Biomass formation in a biofilter begins during the start-up period. During this period,
biomass growth has been described to occur in discrete discontinuous layers, as microcolonies,
that partially cover the solid particles, and that are separated by pores and channels (Shareefdeen
and Baltzis, 1994; Bishop, 1997). Over time bare areas may get covered with biomass. These

layers may consist of cells entrapped within a gelatinous matrix of EPS,produced directly from
the surface-associated microorganisms (Bryers, 1987). The growth of a biofilm is the result of

mass transfer of substrate and its conversion, and it is a function of the transport rate of the
limiting substrate, the biomass yield and/or the biofilm density (Van Loosdrecht et al., 1995).
Heterogeneity in the composition, structure, and distribution of biofilms is a recognised
property of these systems (Bishop, 1997). In a biotrickling filter packed with propylene Pal1
rings and treating ethyl acetate and toluene, density differences were observed between the inner

and outer regions of the biofilm (Schanduve et al., 1996). In general, ce11 density was lower in

the inner region where fragments of lysed cells where observed, even during the acclimation
period. Decreasing concentrations of polysaccharides and proteins dong the biofilm depth have
been reported in aerobic heterotrophic biofilms (Zhang et al., 1997). The base of a biofilm is
considered to be more stmctured and finner than its surface film, which may have an irregular

Ejects of Biomass Growth on Pressure Drop in Bio#fters

Theoretical Background and Lirerature Review

16

topography. In some cases, the biofilm may have the characteristics of a surface film, especially
when filamentous microorganims and protozoa are dominant (Widerer and Characklis, 1989).
The density of a biofilm depends on the type of organisms present, but it also seems to
depend on the fluid dynamic conditions in the biofilter. For instance, higher biofilm densities
have been reported with increasing shear stress on the biofilm (VanLoosdrecht et al., 1995).
Standefer and Van Lith (1993) stated that in biofilters it is possible that there is no
sludge generation if the irnmobilised microbial growth rate, afier adaptation, reaches
equilibrium with the death rate. So dead ce11 matter becomes a partial nutnent source for the
living bacteria, and consequently, sludge does not clog the bed. Nevertheless, this is not always
the case like during the exponential microbial growth phase, which is present during the
acclimation period of a biofilter. Moreover, biofilters can be operated with an exponential
growth rate if the pollutant loading rate is gradually increased during the period of operation
(Cherry and Thompson, 1997).
Even though most studies conceming biofilm structure and characteristics have been
performed with water submerged biofilms, their results are valuable when trying to describe
biofilms exposed to air under moist conditions. Famgia (1999) investigated the development
and properties of biofilms in biofilters; however, more studies of air exposed biofilms. as in the
case of biofilters, are still required.
2.3.2 Factors Affecting Biomass Growth in a Filter Packed Bed

2.3.2.1 Support Material

In general, biomass growth and activity depend on the environmental conditions under
which the microbial cornmunity develops. The physical and chernical properties of a filter's
medium such as water content, superficiai area and porosity, minera1 and nutnent content,
organic content, pH and temperature, significantly influence the characteristics of the
microorganisms that may develop on it, especiaily during the early stages of biofilm

accumulation. Surface roughness has been reported to enhance biofilm development (Van
Loosdrecht et al., 1995). For instance, Hodge and Devinny (1 995) working on the biofiltration
of ethanol vapours with biofilters packed with compost, GAC, and a mixture of compost with
diatomaceous earth, concluded tbat compost supported microorganisms with the highest
Efects of Biomass Growfhon Pressure Drop in BioMers

Theoretical Backaround and Literature Review

17

biodegradation rate constant, suggesting that compost provides a better environment for
microbial growth and activity. In addition, they concluded that GAC particles allow biological
growth only on the surface and in pores sufficiently large for microbial cells. The inner portion
of the particle, which is a substantial portion of its volume, is inaccessible. This observation
could be extended to other inert porous materials with similar characteristics. It has been
suggested that biofilm adherence to a porous support material, such as GAC, is influenced by
the adsorption of the degrading compound on the medium, since the high concentrations of the
adsorbed substrate stimulates biomass growth on the medium's superficial pores (Govind et al.,
1993). Initial cellular adhesion to a specific substratum can be regulated by non-specific classes

of adhesion such as ionic, dipolar, hydrogen bonding, or hydrophobie interactions, and by


specific adhesion mechanisms which require stereochemical or chemical receptor binding
(Bryers, 1987).
2.3.2.2 Substrate Supply

The VOC concentration in the air Stream and the time of exposure to the hydrocarbon
also seem to influence the microbial metabolism that leads to biomass growth, and the biomass
characteristics. Leddy et al. (1995) (Jones et al., 1997) found that P. putida 54G growing on
toluene vapour fonned variant cells that could not degrade toluene but that continued growing in
its presence, metabolising other carbon sources such as organic compounds leaking fiom other
cells or toluene degradation intermediates. Arcangeli and Arvin (1992) (Jones et al., 1997)
reported that the active fraction of a biofilm that was exposed to toluene over an extended
period was only 5% of the total, while nearly al1 of the cells were active during the initial
attachment phase. These observations, obtained under conditions referred as to bacterial injury

or physiological stress (Jones et al., 1997), may have an impact on the biomass characteristics.

Besides the pollutant substrate, which acts as a carbon and energy source for
chemoheterotrophs, microorganisms also require nutrients such as nitrogen, phosphoms,
sulphur, and trace elements for their metabolism. Nutrient addition to biofilters, besides those
provided by the natural medium, has been related to higher elimination capacities since higher
pollutant concentrations are rernoved under the same operating conditions, hence reducing the

Effects of Blomass Growth on Pressure Drop in Biofiters

Theoretical Background and Literature Review

18

biofilter area and cost installation (Wani et al., 1997; Morgenroth et al., 1996). However, the
implications that nutrient addition might have on biomass growth and accumulation has to be
taken into account since an excess of nutrients, given an unlimited substrate, might induce
excessive biomass accumulation and high pressure &op across a biofilter.

For example,

spraying a nutnent solution on a peat bed was reported to cause accelerated biomass growth
near the area of nutrient addition, with final clogging (Govind et al., 1993). On the other hand.
low nutnent availability fiom the natural medium may slow down microbial growth and
distribution, which has detrimental effects on a biofilter's removal capacity.

The source of nutrients has an impact on the arnount of biomass yield. Smith et al.
(1996) studied the effect of two different nitrogen sources, nitrate (NO3-N)
and ammonia (NH3-

N), on the biomass yield in 2 biotrickling filters packed with diatomaceous earth pellets and
used to treat toluene. They used equivalent mass ratios of COD to N, and a nitrogen to
phosphorus ratio of 4. They found that in the biofilter whose nitrogen source was the oxidised
fonn (NO3-N), less nitmgen was utilised for ce11 growth for a given arnount of COD consumed
than in the biofilter whose nitrogen source was the reduced form WH3-N). Based on the

assumption that aerobic heterothrophs accounted for al1 of the net nitrogen utilisation in the
biotrickling filters, it was estirnated that at least 70 % more COD could be degraded for a given
mass of nitrogen in the form of NOJ-N, and at least 40% less biomass, as VSS, is generated for
a given mass of COD using NO3-N as a nitrogen source (Smith et al., 1996). For heterotrophic
microorganisms the growth yield is lower when utilising NO3-N rather than NH3-N as the
source of nitrogen since energy is required to convert nitrate to ammonia for ce11 synthesis.
Hence, Smith, et al. (1996) hypothesised that in the NH3-N fed biotrickling filter, more energy
was available for growth and for the production of exopolysacharides, which supported a large

population of non-toluene degrading heterotrophs.

In an environment high in total salts, the ionic strength of a biofilm may be high, hence
the microorganisms may be expending a significant amount of energy maintainhg the necessary
ionic gradient between the interior and the exterior of the cell, and thus, not producing excess
biomass W e y et al., 1996).
Weber et al. (1994) and Holubar et al. (1995) (Smith et al., 1996) have been successhl in
reducing plugging in trickle bed air biofilters used to treat toluene and a mixture of

Effects of Biomass Growth on Pressure Drop in Bioflters

Theoretical Background und Literature Review

19

hydrocarbons, by using nutrient nitrogen limitation and high ionic strength (NaCl), and both
nitrogen and potassium limitation. However, the VOC removal efficiencies reported were
below 50 %. Schonduve et al. (1 996) observed 50% less biomass growth when nitrate was used
as nitrogen source instead of ammonium in a biotrickling filter used to treat ethyl acetate and
toluene. However, a 70% reduction in the degradation rate was registered.
It seems that nutrient limitation decreases biomass growth because of a decrease in the
degradation rate. Moreover, some studies show that the use of nitrate as a nitrogen source may
reduce biomass growth without afTecting the removal rate; some others, on the contrary, indicate
a reduction of both microbial growth and degradation capacity. More research is needed to
elucidate these contradictory results.
2.3.2.4 Other Factors

Figure 2.1 presents a schematic of several factors influencing biofilm structure and
accumulation. Some of them have been discussed in the previous sections, others are briefly
outlined below.
While at high substrate loading rates and low shear stress a more heterogeneous biofilrn
may form, at low loading rates and high shear stress a smoother, but patchy in distribution
biofilm may develop. Flow shear stress rnight not apply to biofilters because of its low values;
however in aquatic systems it is relevant. The types of microorganisms also affect the structure

of the biofilm: fast growing microorganisms tend to fom a weaker and more porous biofilm
than slow growing ones (Van Loosdrecht et al., 1995). It is believed that if a smooth biofilm

develops, then its thickness will depend on the substrate loading rate. Furthemore, in reactors

with low shear stress, thick and highly heterogeneous biofilms will tend to form.

Effets of Biomars Growth on Pressure Drop in Bioffters

20

Theoretical Backwound and Literuture Revie w

bstrate surface
loading rate
& nutrients

T y p e o f m i c r o o r g a n ism s
I

Y ield c o e f f i c i e n t

A m o u n t o f biofilm

T y p e o f reactor

nsity

1 S hear rate (

BIOFILM S T R U C T U R E A N D A C C U M U L A T I O N

Figure 2.1
Relation of factors afTecting biofilm growth and its accumulation (Van
Loosdrecht et al., 1995).

2.3.3 Biomass Distribution in the Packed Bed of a Filter


A microbial stratification in type and number dong a biofilter bed is expected. Several

researchers have investigated the spatial distribution of microorganisms in biofilters, and have
observed that the density of microorganisms is greater at the inlet section where readily
degradable substrates and nutrients are available and more removal takes place (Corsi and Seed,
1995). Kinney et al. (1996) reported four orders of magnitude greater numbers of platable

microbes in the inlet section than in the outlet section in a biofilter packed with Celite porous
silicate pellets that treated toluene. Kosky and Neff (1988) (Swanson and Loehr, 1997)
observed that deeper into a biofilter bed, smdler populations of different organisms existed,
probably adapted to low concentrations of a more complex substrate.

Similar biomass

distribution was observed by Mohseni and Allen (1997), who reported that more than 60 to 80%
of the total pressure &op in biofilters packed with compost, wood chips, and perlite or GAC

were measured in the upper sections of the biofilters, which where operated in a down-flow

mode. They explahed this behaviour as a result of higher substrate concentrations and higher
Effectsof Biomuss Growth on Pressure Drop in Biofiters

Theoretical Background and Literature Review

21

microbial activity in the fwst sections of the biofilter, which resulted in more biofilm growth,
less void space and greater pressure drop. An altemating feed strategy consisting of switching
the flow direction fiom inlet to outlet, and viceversa, allowed for less clogging and a more
unifomly distnbuted biomass profile across the length of a biofilter treating toluene vapours
(Kinney et al., 1996). A similar advantage of reduced biofilm thickness near the inlet region is
achieved in unidirectional flow biofilters subject to discontinuous feeding (Wright et al.,1998).
These observations suggest that biomass concentration in a biofilter packing material is
proportional to removal rate. Pedersen and Arvin (1997) concluded that a constant toluene
liquid concentration through a biotrickling colurnn caused an even biofilm growth along the
filter height. They measured a constant average value of biofilm thickness and a constant
average value of content of active biomass along the column. Cherry and Thompson (1997)
have suggested that during exponential growth in a biofilter, biomass generation is proportional
to the conversion of substrate, hexane in their experimental case. The proportionality constant

is a yield coefficient, which relates the conversion of hexane and the formation of biomass.
On the other hand, studies with a biotrickling filter packed with Pall rings and used to
treat ethyl acetate and toluene, showed that the increase in biofilm thickness did not correlate
with an increase in degradation of the pollutants (Schonduve et al., 1996). It is believed that an
increase in biofilm height can be linked to a decrease in the portion of the biologically active
layer of the biofilm due to nutrient and oxygen limitations. This type of observation, however,

has been scarce.

More recently, Deshusses and Cox (1998) by means of Computer Axial Tomography
(CAT) studied the packed bed of both a biotrickiing and a biofilter treating toluene vapours.
The CAT scans of the biofilter packed with a mixture of wood chips and compost (80120 by
volume), as well as of the biotrickling filter packed with Pall rings, revealed that biomass grew
very heterogeneously. Some parts of the medium were covered with a thin biofilm, whereas
other regions were completely clogged, especially for the case of the biotrickling filter with high
biomass content. In the biofilter medium, a heterogeneous biofilm was observed even at the sub
millimetre scale. The observed microscopie roughness of the biofilm dong with previous
observations of channels in biofilms conducted in another study using confocal laser

Effects of Biontass Growth on Pressure Drop in Biofiters

22

Theoretical Background and Literature Review

microscopy, support the hypothesis that biofilm roughness contributes to a significant extent to
the increasing interfacial area for polllutant mass transfer (Deshusses and Cox, 1998).
Kinney et al. (1996) used a biofilter packed with Celite porous silicate pellets to treat
toluene, and they observed that in the first 25 cm of the bed (first section), 75% of the inlet
toluene was degraded during the first 20 days of operation. However, after day 20, the removal
eficiency dropped off rapidly. Since the bed was remixed and the removal efficiency continued
to decrease not only in the first section but also in the second section, they related this decrease
in eficiency to the way biomass accumulated on the pellets. Kimey et al. ( 1996) hypothesised
that the microorganisms colonised both the inner pores and the surface area of the porous
pellets. As the biofilm thickened in the first section due to high toluene degradation, the
openings to the micropores clogged, toluene becarne unavailable, and the microorganisms in the
inner pores became inactive.

Consequently the overall microbially active surface area

decreased, diminishing toluene mass transfer and degradation.

Using scanning electron

microscopy, biomass penetration into the pellets was detected to be not beyond 100 Fm.

2.3.4 Effects of Biomass Growth on the Flow Characteristics


Several authors (Soria1 et al., 1997) have studied pressure drop due to excess biomass in
biotrickling filters, and possible biomass accumulation control strategies. This is not the case
for biofilters, where the interacting phases are gas and biofilm, and biomass growth patterns and
biomass accumulation control strategies suitable for water based biotrickling filters are
different. Although the influence of biomass accumulation on the flow characteristics is more
significant in biotrickling filters that treat odors, VOC's and toxic compounds, similar problems
can be encountered in biofilters and analogous effects by biomass growth on the airfiow
charactenstics can be expected. Biomass accumulation in porous media can substantially
reduce the capacity for transport of mass and momenturn within a biofilter bed. Biofilm growth
on the surface of the filter medium hinders airflow through the bed and reduces the effective
pore space of the bed, and this reduced effective pore space results in decreased specific
permeability and an increased fiction factor. Several studies in the areas of groundwater
bioremediation and petroleum recovery have reported decreased water permeability due to

Effects of Biomass Growth on Pressure Drop in Biofiters

Theoretical Background and Literature Review

23

biofilm accumulation in the effective pore space, where EPS seemed to play an important role in
plugging (Cunningham et al., 1990).
Furthemore, it has to be considered that if biofilm thickness remains small cornpared to
the effective pore space, accumulation of biofilm will not affect the distribution of pore
velocities within a porous medium, as pointed out by Cunningham et al. (1990). However, if
biofilm occupies a significant fraction of the effective pore space, decreases in effective porosity
and specific permeability will occur.

It is well established that over time channels may fonn in a biofilter bed packed with
natural materials. This phenornenon has been related to insuficient or excessive moisture
content or aging of the medium, which causes the bed to crack, to reduce its volume, and to
compact.

Biomass growth, however, may also play an important role changing the

characteristics of the bed, and ultimately the airflow charactenstics through a biofilter. In recent
studies using Computer Axial Tomography, the presence of small and large channels, from less

than 5 mm2to 380 mm2,caused by biomass accumulation, was detected in a biotricWing filter
packed with Pal1 rings. Similar channelling was observed in a biofilter packed with compost

and wood chips, yet in this case most of the air channels were detected on the side, where wall
effects might have occurred (Deshusses and Cox, 1998).

In a biofilter packed with Celite porous silicate pellets used to treat toluene, overgrowth
of biomass in portions between the packed pellets clogged the void space, resulting in air
channelling and a rapid decrease in removal efficiency (Kimey et al., 1996). Interestingly, no
increase in pressure drop was observed, which might indicate that channelling not always causes
a pressure drop build-up, probably due to the heterogeneity and tortuosity of the channels
fonned.
Deviation of the airflow due to biomass clogging in certain parts of the filter bed disnipts
the homogeneous airfiow through the bed. This results in locally higher air permeabilities, and
since more air is passing through a smaller cross sectional area, the pressure drop increases and
the elirnination capacity of the biofilter decreases.

Theoretical Backmound and Literature Review

24

2.3.5 Biomass Growth Quantification in Packed Beds


The procedures for measuring biomass accumulation in packed beds, also referred as
biofilm growth, can be divided into two main techniques (Characklis, 1990): direct and indirect.
Direct measurement of the quantity of biofilm can be conducted by weighing the dry or wet
biofilm mass accumulated on the support medium, or by detemining the wet biofilm thickness
and its distribution by microscopie techniques such as scanning electron microscopy,

transmission electron microscopy, and confocal scanning laser microscopy (Fletcher, 1992:
Stewart et al., 1995). Indirect techniques are classified based on the particular constituent or
property of the biofilm fiom which the amount of biomass is inferred. Table 2.2 shows some of
these techniques.
Some indirect techniques for biofilm and biomass quantification (Adapted from
Table 2.2
Characklis (1 990) and Tate III (1 995))

1 Anahtical
. Techniaue

1 Classification

Specific biofilm constituent

Microbial activity within the biofilm

1
Effects of biofilm on transport properties

--

Polysaccharides (chitin in fungal cells/


muramic acid in bacterial ceils)
Total organic carbon
Chernical oxygen demand
Protein
Lipids
Cellular carbon by chloroform fumigation
Viable ce11 count
Epifluorescence microscopy
Adenosine Triphosphate (ATP)
Substrate removal rate
1
Respirometry (Oz consumption and CO2
production)
Friction resistance
Heat transfer resistance
1

Even though several of these techniques have been used in biofilters and biotrickling
filters to characterise and measure biomass concentration in the filter bed, their application has
to be evaluated for each particular case since each technique presents advantages and
disadvantages (Singh et al., 1994).
Govind et al. (1993) measured CO2 production in biofilters treating mixtures of VOC's,
packed separately with peatkompost, pelletised activated carbon, and ceramic Celite.
-

---

Effects of Biomuss Growth on Pressure Drop in BioJiters

Theoretical Background and Literaure Review

23

Cumulative CO2over time was measured, and also calculated assuming that al1 the compounds
removed were converted to CO2. Based on the comparison of calculated and measured CO2
production, conclusions related to the carbon balance, specifically to biomass generation, could
be made. Kinney et al. (1996) used the ratio of CO2evolved across a biofilter column over the

toluene degraded in CO2 equivalents to represent the fiaction of carbon from toluene
degradation that was released from the column as CO2. The remaining carbon was converted to
biomass considering a complete degradation of toluene occurred. This is an indirect way of
measuring biomass growth, which relies on the stoichiometry of the mineralisation of substrates
and ce11 production.

Although with this method the amount of biomass growth can be

estimated, it does not provide information about biomass distribution or its growth patterns. In
addition, this respiration technique, as with other microbial activity related procedures,
quantifies only active biomass, which is a portion of the total biofilm. As a result, these
techniques are disadvantageous when trying to assess the influence of biomass accumulation on
pressure drop in packed beds since not only active biomass but also its extracellular substances
occupy the bed void space.
Several other techniques have been used in biofilters and biotrickling filters. Biofilm
thickness, measured as an average thickness by weight of the total arnount of biomass on the
bed sample, was used to quanti@ biomass growth in a biotrickling filter treating toluene and
packed with polyvinyl difluoride cubes (Pedersen and Arvin, 1997). Bacterial enurneration in
plate count agar has been applied for an industrial biofilter packed with popiar wood bark and
treating toluene vapours (Andreoni et al., 1997), and dso in biotrickling filters packed with
diatomaceous earth pellets used to treat toluene (Smith et al., 1996). The results of the number
of bacteria have been expressed as most probable number (MPN) per dry weight of medium or

as colony forming units (CFU) per dry or wet weight of medium. Measurement of active
biomass in biofilters has been accomplished indirectly by determining cellular ATP (adenosine
triphosphate) content in biomass samples (Andreoni, et al., 1997), and the amount of the protein
in disintegrated biofilm samples fiom a biotrickling filter treating toluene and packed with
polyvinyl difluoride cubes (Pedersen and Arvin, 1997). For the case of the protein analyses,
since the ce11 protein and extracellular substances protein contents may vary, assays based solely
on protein may underestimate the amount of total biomass. On the other hand, overestimation
Eflects of Biomass Growth on Pressure Drop in BiofiIters

Theoretical Background and Literuture Revie w

26

of the active biomass may happen if there is a large pool of protein accumulation due to cell
injuries (Jones et al., 1997). Errors due to composition variation of cells and extracellular

substances may arise using the polysaccharide technique.


Estimation of microbial biomass using direct techniques seems to give a better
description of b o l the amount of biomass accumulated on the bed and its growth pattern.
Nevertheless, the application of microscopy is restricted to relatively flat surfaces and cannot be
applied to porous or topographically complex materials (Fletcher, 1992). The determination of
microbial mass using wet weight has been considered imprecise because intracellular water, ce11
wetting water, and water contained in extracellular substances are also taken into account, and
the real microbial mass can be then overestimated (Singh et al., 1994). This is true when just

protoplasm has to be quantified; however, when the density of the biomass is relevant, as for
pressure drop effects, wet weight may actually describe the slimy and capsule forming
characteristics of the biomass, given by the extracellular polymea.
The use of rnicroscopic techniques for analysing biofilm characteristics has shown the
advantages of these methods for studying structural heterogeneity in biofilms, as well as for
establishing biofilm coverage and thickness (Stewart et al., 1995; Silyn-Roberts and Lewis,
1997), which is relevant for understanding biomass effects on flow characteristics through

biofilm systems. Confocal scanning laser microscopy seems to be superior than the other
techniques, since it enables one to visualise the intemal structure of the biofilm and allows a
three dimensional analysis; however it may not be applicable to thick biofilms and it requires
expensive equipment.

2.4 Modelling of Biomass Growth and Its Effects on Pressure Drop


2.4.1 Modelling Biomau Growth in B i o l t e n
Modelling of biomass growth and its distribution on the packing material have been
considered when denving predictive models of biodegradation in biofilters.

Several

assumptions related to biomass growth have been taken at a rnicroscopic level to simpliQ the
rnodels. Some assumptions of this kind have been considered in the derivation of a quasisteady-state mode1 based on the growth of biomass for describing methanol biofiltration
(Shareefdeen et al., 1993; Shareefdeen and Baltns, 1994):

Eflects of Biomass Growth on Pressure Drop in Biojiters

Theoretical Background und Literature Review

27

1. n i e biolayer is formed on the exterior surface of the particles and it does not grow in

the pores of the particles and thus, no reaction occurs in the pores.
2. The biolayer does not form uniformly around particles; there may be patches of bare

particle surface.
3. The biofilm density, defined as the amount of dry biomass per unit volume of

biolayer, is constant. Also considered by Deshusses er al. (1995).


4. The biolayer surface area per unit volume of reactor is constant dong the column.

5. Biomass does not accumulate in the filter bed and thus, the specific biolayer surface

area is constant.
6. The effective biolayer thickness varies dong the biofilter column. By solving the

mass balance equations of the model, the values of the effective biolayer thickness along
the bed were estimated, and a linear correlation between biofilm thickness and the
compound concentration in the air at a specific position along the biofilter was found.
7. Since the fraction of the extemal surface area of particles covered with biofilm is hard

to estimate, the value used in the model was estimated as the quotient of the volume of
inoculum suspension used and the volume of initial packing material. This neglects the
amount of biomass accumulated dunng the start-up period.
The quasi-steady-state model developed by Shareefdeen and Baltzis ( 1994) includes

consideration of real biomass growth patterns obsewed in biofilters, such as assumptions 2 and
5 above, but aiso includes assumptions that are less realistic, but help to simplify the model and

allow for the prediction of the biodegradation process. Shareefdeen et al. (1993) recognises that
the biofilm density in a biofilter needs to be deterrnined and that biofilm density decreases as the
thickness of the biofilm increases. The biolayer surface area per unit volume of reactor also
should increase as the biolayer thickness decreases towards the exit of the biofilter. Cherry and

Thompson (1997) pointed out that the assumption that there is no net biomass growth in
biofilters, as in some cases has been observed, implies that biofilters should be modelled
c o n s i d e ~ ga maintenance metabolism or the equivalent situation of balanced ce11 growth and
death. However, this has not been the case in most of the models for biodegradation in
biofilters. Cherry and Thompson (1997) emphasised that substrate consumption following firstorder kinetics or some form of Monod kinetics coupled with the assumption of steady-state for
Effects of Biomms Growth on Pressure Drop in Biofifers

Theoretical Background and Literawe Review

28

ce11 growth or no net ce11 growth are inconsistent, and do not allow to model biofiltration during
the acclimation period where exponential biomass growth is more likely to happen.
Cherry and Thompson (1997) proposed two ways of estimating the biomass content in a
biofilter during exponential growth. The first considers the amount of substrate converted and
uses the yield of cells fiom substrate during active growth to estimate the total biomass created.
The second approach uses an apparent first-order rate constant for a plug 80w reactor given by:

where kl is the apparent first order rate constant, T is the residence time in the bed, and Sou,and
Sin are the outlet and inlet substrate concentration in the gas.

Alonso et al. (1997) developing a model to describe physical and biological processes
taking place in a biotrickling filter for waste gas treatment, proposed the following expression
for calculating biomass-affected porosity (sf) in a bed packed with spheres, where biomass
grows uniformly on the spheres' surface:

where E* is the initial porosity of the bed without biofilm, LJ is the biofilm thickness, R is the
radius of the sphere equivalent to the packing medium, n is the number of packing spheres in
contact with a single sphere or coordination nurnber. Equation 4 considers that the spheres that
were in contact when the biofiiter was clean will remain in contact, and therefore the biofilm
will grow only on the void space left between the solids (Alonso et al., 1997). Another
expression for predicting bed porosity with biofilm growth was proposed by Cunningham et al.

(1991). This expression was deduced for the case of uniform porous spheres packed in a cubic
arrangement, where 8 spheres are considered to fit in a cube of sides equal to two times the
diameter of the spheres. The proposed expression is:

where dis the sphere diameter.

Effects of Biomass Growth on Pressure Drop in Bioflters

iVieoretical Backmound and Literature Revie w

29

2.43 Modelling the Effects of Biomass Accumulation on Pressure Drop


2.4.2.1 Models Developed for Groundwater Flow in Porous Media
Microbial growth in porous media as a cause of specific permeability reduction and
increased pressure dmp is a phenornenon that has been studied in the areas of groundwater
systems, soi1 science, water treatment, and petroleum engineering (Cunningham a al., 1990).
Water specific pemeability reduction of 65 to 99% due to biofilm accumulation has been
reported in studies that have been conducted using a variety of porous media, including core
samples fiom field sites as well as synthetic porous media (e. g., glas spheres), and both pure
and mixed cultures as inoculum. Nevertheless, Taylor and J& (1990) have stated that these

have been just descriptive atternpts for understanding the process and there have been only few
studies that have tried to link the dynamics of biofilm growth and substrate utilisation to water
specific permeability reduction and pressure drop increase. Some of them are cited in the
following paragraphs.
Crawford (1987) (Cunningham et al., 1990) carried out laboratory experiments with a
capillary reactor containing a single layer of 1 mm qlass spheres in order to demonstrate the
relationship between biofilm accumulation and the hydraulic resistance of a porous medium. It
was observed that after the biofilm thickness reached a steady value, the friction factor still

increased, suggesting that the biofilm surface continued to develop in a progressively irregular

manner. Even though an increase in the fi-iction factor due to biomass accumulation occurred,
the presence of biofilm did not alter the observed relationship between fiction factor and
Reynolds number for flow through porous media (Cunningham, et al., 1990).
In another study, Taylor and Jaff (1990) perfonned experiments to quanti@ the water
specific permeability reduction caused by enhanced microbial growth in columns packed with

sand, which was used as a typical porous medium. From piezometric head measurements at
different levels of the bed, the hydraulic conductivity was detennined for each sample port using
a unidirectional form of Darcy's law

Effects of Biomass Growth on Pressure Drop in BiuJlters

30

Theoretical Background and Literature Review


where Q is water flow rate (L.'/'T'),

A is the cross sectional area ( L ~ ) ,K is the hydraulic

conductivity (Ln), and dW<Ix is the variation of piezometric head in a distance x along the
column A general expression of Eq. 6 is

and AP is the pressure drop along a height L in the


where U is the superficial velocity 0,
column. With the hydraulic conductivity K,the specific permeability k was calculated fiom the
definition of hydraulic conductivity:

where k is the specific permeability (L~),p is the fluid density

g is the acceleration of

gravity (UT*),
and p is the fluid viscosity (M/LT). In this way, the reduction in specific
permeability was estimated as the ratio of actual specific permeability over its initial value. For
the sand columns under aerobic conditions, Taylor and Jaff (1990) obtained an expression of
specific permeability reduction, expressed as the fraction of the initial specific permeability, as a
function of biomass concentration in the bed.

Even though this expression cannot be

generalised, it is an attempt to describe specific pemeability changes due to biomass


accumulation. Moreover, they identified the need to obtain an expression to correlate biomass
and specific permeability reduction, but based on a physical model.

The hydraulic conductivity, K,is a function of the porous medium properties and of the
hydraulic properties, and provides a measure of the ability of the porous medium to conduct
water. It is ofien useful to express the specific pemeability of a porous medium as a property of
the medium independent of the density and viscosity of the fiuid. Indeed, the definition of
specific permeability as a huiction of hydraulic conductivity and the fluid properties allows the
employment of the hydraulic conductivity in systems where the fluid has a viscosity and a
density different fiom water. In addition, specific permeability is an intrinsic property of the
porous medium, and depends on pore size distribution, pore shape, tortuosity, specific surface,
and porosity (Taylor et al., 1990).

In a later study, Taylor et al. (1990) obtained expressions to analytically predict changes

in porosity and specific surface in order to estimate specific penneability changes caused by
Effects of Biomass Growfhon Pressure Drop in Biojilters

31

Theoretical Backmound and Literaure Review

biofilm growth in a rigid porous medium. For developing the model, some assumptions were
taken into account: the rigid porous medium is saturated with a single fluid, Darcian flow
(lamina. fiow) predominates, the biofilm has a negligible specific permeability, and aerobic
substrate utilisation conditions prevail.

They took as a basis the single-phase specific pemeability Kozeny-Carman equation,


which considers that the porous medium is equivalent to a conduit, whose cross-sectional area
has a complicated shape but a constant area. The conduit diarneter was taken to be 4 times the
hydraulic radius, defined as the flow cross-sectional area divided by the wetted perimeter
(Taylor et al., 1990):

where k-is specific permeability,

is a constant equal to 115, n is the porosity, and M is the

specific surface (L2A,3). Two modelling approaches were considered.


a) Mode1 deduced for a regular packing medium of uniform spheres. From geometric
considerations, Taylor et al. ( 1990) obtained

n=l-[&J
,and (Eq. 10)

where a, is the packing arrangement factor, with the subscript m indicating number of contact
points and d the sphere diarneter. For a6 (cubic arrangement)=l, ad (orthorhombic)= 3 '"/2, ai*

(tetragonal-spheroidal)=3/4,and a 12 (rhombohedral)= 2?
If the variation of n and M with biofilm thickness for a given packaging arrangement is
specified, then the changes in specific permeability due to changes in biofilm thickness can be
detemiined. Expressions for porosity (nb) and specific surface (Mb)with biofilm thickness for a
heterogeneous size distribution of spheres were calculated by (Taylor et al., 1990):

E#ects of Biomass Growth on Pressure Drop in Biojilters

Theoretical Background and Literafure Review

32

where d is the spheres' diameter and Lf is the biofilm thickness. The biofilm-affected specific
permeability was calculated by substituting nb and

Mb

in the Kozeny-Carman relation (Eq. 9),

assuming co remained constant with biomass growth, and that al1 spheres were coated with an
impermeable film of biomass of constant thickness.

b) Cut-and-random-rejoin-type model for porous medium with pores of various radii


randomly distributed in space. This approach includes specific permeability statistical models
that consider the random nature of the intercomectedness of the pores.
Mualem (1976)(Taylor et al., 1990) proposed a general specific permeability model that
considers saturated moisture content. Mualem obtained

where ~(~,r,p'),
is the correction factor accounting for eccentricity of the flow path (tortuosity
is a correction accounting for partial correlation between the pore radii r and
factor), G(~,r,p')
p' at a given moisture content 8(r). f(r) is the pore size distribution function, r and

are the

pore radii of a pair of serially connected capillary elements, ro is the minimum pore radius, and

R the maximum pore radius.


It was assurned that the tortuosity and correlation factors are power functions of
porosity:

where r and y are dimensionless constants. Finally specific permeability was given as

(Eq*16)

Moisture content, porosity, and specific siirface were also expressed in tenns off(r):
r-

- -

--

--

--

Efects of Biornass Growth on Pressure Drop in Biofilters

33

Theoretical Background and Literature Review


R

f (r)dr

n=

,and (Eq. 18)

ru

"

rn

(Eq. 19)

In this case the changes in specific permeability due to changes in biofilm thickness are
determined if the variation off(r) with biofilm thickness is specified.
For estimating the biofilm-aHected specific pemeability, nb and Mb were also
determined in terms offb(r), which is the associated pore size distribution describing the void
space not filled with biomass, and is a function of the biofilm thickness. Taylor et al. (1990)
obtained expressions to estimate specific permeability with and without biofilm growth using
Mualem's specific permeability model, limiting their analysis to the case where f(r) was given
by a power function. Finally, they found an equation to predict specific permeability reduction
due to biomass growth in terms of several variables such as biofilm thickness, minimum and

maximum pore radii, and a pore size distribution index.


Computational results fiom these equations showed that specific pemeability reduction
is greater when the medium has uniform pores sizes, and the application of the sphere model is

limited to relatively thin biofilms on homogeneous, unconsolidated media. It was noted that the
cut-and-random-rejoin models are superior to the sphere model because they can be applied to
porous media with a wide range of pore sizes, and they exhibit the proper asymptotic behaviour

es the pore space fills with biomass (Taylor, et ai., 1990). It is important to underline that a
major assumption made in this analysis is that of a uniform biofilm thickness. In order to
descnbe the variability of the biofilm thickness, a specific permeability model could be
manipulated to yield expected values of specific permeability with biofilm growth, if a
probability distribution of biofilm growth were available.
Even though the Mualem-based specific pemeability model and porosity model c m be
used to make reasonable predictions of these parameters for a porous media affected by biofilm
(Taylor and J e , 1990), they are complex models requiring values of parameters not easily
available. Besides, the biofilm approach used assumes continuous and uniform biomass growth
on the exposed surface of each particle in the porous medium, and as stated by Bayeve and
Valocchi (1989) (Clement et al., 1996), these authors did not support evidence to ver@ the
Efects of Biomass Growth on Pressure Drop in Bioffters

Theoretical Background and Literature Review

34

growth patterns, and hence the assumptions taken. Furthemore, it is more probable that
biomass grows in a heterogeneous fashion as stated in previous sections of this work.
In a more recent study, Clement et al. (1996) developed analytical equations to model
changes in porosity, surface area, and specific permeability caused by biomass accumulation in
porous media exposed to water flow. nieir analysis was based on macroscopic estimates of
average biomass concentrations and it did not assume any specific pattern for microbial growth.
From the three different approaches to model microbial growth and accumulation processes in
porous media: continuous biofilm (model by Taylor et al. (1990)), discrete micro-colony, and
macroscopic approach, they developed the macroscopic approach, which considers only
spatially averaged biomass concentration as the model variable.

On the other hand, the micro-colony model (Vandevivere and Baveye (1992), Clement
et al. (1996)) considers discrete microbial colonies or sparse rnicrobial growth. In reality,
biomass growth may involve both continuous and patchy growth patterns, hence a macroscopic
approach considering spatially distributed averaged concentrations seems to be more realistic
(Clement et al., 1996).
The main assumptions taken by (Clement et al., 1996) in order to develop their model
are:

1. The Mualem's cut-and random-rejoin model (Taylor et al., 1990) is applicable.

2. The pore-size distribution function can be derived fiom soil-water (drainage)


retention functions, as the van Genuchten and Brooks and Corey functions. It remains
constant with biomass growth.

3. Microorganisms are assumed to exist in both aqueous and solid phases, and biornass
growth and nutrient consumption occur in both phases.
4.

Changes in porous media properties are caused by accumulation of solid-phase

biomass in pore spaces.

The fmai macroscopic mode1 equations, which can be used to compute biomass-afTected
specific permeability, specific surface area, and porosity values at any specific attached biomass
concentration, are:

Effects of Biumass Growth on Pressure Drop in Biofiters

Theoretical Background and Literahve Review

where d i s the biomass volume fraction (volume of biomass/total volume),

35

is the mass (dry

weight) of microbial cells per unit mass of aquifer solids, pk is the bulk density of aquifer solids,
p . is the solid-phase biomass density, no is the unaffected porosity, and nb is the biomass-

afTected porosity.
This model is in agreement with the macroscopic-level Darcy equation fiom which the
models are developed. Furthemore, for the case of a size distribution index equal to 3, the
results fiom the macroscopic model are in exact agreement with the predictions of continuous
biofilm model of Taylor et al. (1990), and the model is computationally more efficient (Clement
et al., 1996).
2.4.2.2 Models Developed for Bioreactors

Other studies have investigated the correlation between biomass growth and pressure
drop in bioreactors, such as in a solid state fermentor and in an aerated cocurrent upflow fixedbed bioreactor,

Experiments carried out with solid state fermentors (SSF) (Auria et a l , 1993) have
revealed that the variation of pressure &op observed as air passes through the packed bed was
linked to mould growth due to reduction in the bed void fraction. In these experiments, an
increased pressure drop dong the column was clearly conelated to an increase in biomass
concentration over time at different airflow rates. Auria et al. (1993) determined that lower
relative permeabilities, k&,

where k is the initial specific permeability and kb is the biofilm-

afTected specific pemeability, correlated to higher biomass concentrations in the bed.


Considering that in most SSF applications a laminar flow regirne has been found to
prevail and kinetic energy losses are negligible, the Ergun equation (Eq. 1) used for fluid flow

through a porous bed was simplified to Darcy's law (Eq. 5). Darcy's law was used to estimate
Effects of Biomass Growth on Pressure Drop in Bioflters

Theoretical Background and Literature Review

36

the hydraulic c~nductivityfiom pressure &op data, and specific pemeability was then
calculated using the definition of hydraulic conductivity (Eq. 8). Finally, a linear correlation
equation was obtained to estimate the mould growth based on the observed decrease in specific
permeability. This expression, suitable for the inert resin used as packing material, is not
suitable for estimating biomass concentrations in supports that are modified during microbial
growth since the changes in specific permeability may not be only attnbuted to biomass
accumulation, but also to macroscopic changes in the suppon such as compaction.
Nevertheless, in experiments where Auria et al. (1993) used natural support media that acted as
substrate source too and underwent changes in their properties, the same behaviour of pressure
drop vs. time was observed as with the inert medium. This suggests that similar phenornena
affecting pressure drop in the inert medium take place in the natural media. However, in this
last case other factors such as water sorption and compaction, should also be considered (Auria
et al., 1995).

In another study, Deront et al. (1998) investigated the possibility of evaluating biomass
accumulation by pressure drop measurements in an aerated concurrent upflow fixed bed
bioreactor, which was continuously fed with wastewater containing industrial organic
pollutants. They obtained an expression that related pressure &op, the Reynolds numbers of the
gas and liquid, and the equivalent diarneter of the particles used as packing material, clay balls:

where A P L is the pressure &op dong a specific height, de is the equivalent diameter of packing
particles, VSG is the superficial velocity of the gas, pc is the gas density, Re is the gas Reynolds
number, and R ~ isL the liquid Reynolds nurnber. The Reynolds nurnbers are defined as:
Re, = V s d , PG

Re, = G d , P L
where p and p denote the density and viscosity of the gas (G) and liquid (L), VsL is the
superficial liquid velocity, and d, is the average particle diameter of the clay balls.

Efjects of Biomass Growth on Pressure Drop in Biofilers

37

Theoretical Background and Literature Review

Using Eq. 24, fiom pressure drop measurements, and the operating conditions, the equivalent
diameter (de)can be calculated. Then, the biomass-affected porosity ( E ~ )CM be estimated from
the expression of equivalent diameter:

Hence, the biofilm thickness (Lj)can be estimated, assuming a planar geometry, from:
E* =go

-a,L,
Y

(Eq*28)

where EO is the initial porosity, a, the specific area of the packed bed, and both parameten are

known fiomthe packing materiai specifications. This approach used by Deront et al. (1998)
applies for two-phase flow through a packed column.
Deront et al. (1998) also measure volatile suspended solids fiom suspended biomass
detached fiom fixed biomass fiom the substratum. These values divided by the estimated
biofilm thickness gave a biofilm density. Consequently, an expression correlating biofilrn
density and biofilm thickness was obtained. In this way, based on pressure drop measurements.
biomass growth and accumulation could be estimated.
Modelling the relationship of pressure drop and biomass accumulation in these
bioreactors has been done in order to be able to estimate the amount of biomass growing in the
packing meterial based on physical operating pararneters such as pressure drop. Nevertheless,
their approaches are valuable when trying to mode1 the effects that biomass growth has on
pressure drop in biofilters, especially for the case of an SSF, which operates similarly as a
biofilter,
Both the models developed for groundwater flow and for bioreactors take into account
that biomass accumulation in porous media results in the reduction of the interparticle void

space, measured as porosity, and in the change of the hydrodynamic characteristics of the
media. Al1 these models have involved key hydrodynamic variables (Cunningham et a!., 199 1):
porosity describing the fiee pore space, specific permeability descnbing the conductive
properties of the media, and the Giction factor that quantifies fictional resistance. And even

though some of these pararneters are evaluated for water flow, there is the possibility of using

Effects of Biomm Growth on Pressure B o p in Biofhers

Theoretical Background and Literature Revie w

38

them for aimow, given that they can be calcuiated considering air properties such as density and
viscosity.

Effects of Biomass Growth on Pressure Drop in Biojlfers

Eiperimental Procedure

39

3 EXPERIMENTAL PROCEDURE
3.1 General Approach
The experimental approach involved the operation of a biofiltration system to study the

effects of microbial mass growth and distribution on pressure &op, and their correlation during
the removal of methanol. The system was operated for 150 days during which time excessive
biomass accumulation was achieved that was frequently monitored along with pressure drop.
At the end of this period, similar monitoring on biomass development and pressure drop was
performed during the biofiltration of a-pinene for 60 days.
Hardwood chips and inert pellets, called Nova lnertQ, were used as packing materials.
Wood chips were used as a natural packing material that is available on site in pulp and paper
mills where this technology has a potential application. While the feasibility of wood chips
alone as a biofilter packing matenal was assessed, the Nova Inert pellets were used since they
allowed for interference-fiee biomass quantification, and a shape- and size-defined particles for
modelling the effects of biomass accumulation on pressure drop. Nutrients were provided in
excess to both packing materials to ensure enough biomass growth. In the rest of this work,
Nova Inert will be used as a reference to Nova lnertB.
Biomass quantification along the bed and pressure drop measurements were
complemented with tracer tests as a means of assessing the effects of biomass accumulation on

the airflow characteristics in the biofilters. Simultaneous to the study of biomass effects on
pressure drop conducted in this research project, the biofilm characteristics and biofilm
development on the biofilter beds were investigated by two other students (Farrugia, 1999).
Information related to analytical techniques and experimental procedures are presented
in the following sections.

3.2 Experimental Set-up

The biofiltration system consisted of two identical bench-scale biofilters which were
operated in a dom-flow mode to remove frorn a moist air Stream methanol, in a first run,and apinene, in a second run (see Figure 3.1 and 3.2). For each biofilter, building compressed air was
passed through a ~ l e x i ~ l ahumidification
ss~
column (0.1 5 m diameter, 1.55 m height) packed
Effects of Biomass Growth on Pressure Drop in Biojlters

Experimentul Procedure

40

with Intalox saddles (2 cm diameter), to hurnidifi the air to 100% relative humidity and control

its temperature. The air stream temperature was controlled by the humidifier's circulating
water, which allowed the bed temperature in the biofilters to be in the mesophilic range of 35 to
40

OC.

Water in the humidifier was continuously circulated by a pump (model 3E-12N, Little

Giant Pump Co., Oklahoma City, OK), and the water flow rate was controlled manually and its

temperature was maintained by a heater (model 11 12, V W R Scientific, Polyscience, Niles, IL).
A small and controlled airflow fiorn the

main inlet air stream was passed through a stainless

steel vesse1 containing liquid methyl alcohol (Sigma-Aldrich, 99.8 +%, ACS reagent), and
mixed with the main humidified air stream, which then was fed into the biofilter's top. Airflow
rates (flowmeter main Stream: model F 1- 1SO2A, Omega Engineering Inc., Stamford, CT;
polluted Stream: Scott Specialty Gases, Inc., Troy, MI), temperature and pressure in the
biofilters and humidifiers were controlled by a control panel.

Flow
contr

parger

voc

l-

W o o d chips or
N o v a Inert@

Air

W ater bath
Figure 3.1

1 Treated air

Schematic representation of the experirnental set-up for one biofilter

Each biofilter consisted of a plexiglassBcolumn with four removable sections connected

in series. Each section had a diameter of 28 cm and a height of 30 cm. Perforated stainless steel
plates (7 mm mesh diameter) were used as bed supports and as flow distributors in each section.
Effects of Biomass Growth on Pressure Drop in Biojiters

E.perinzentu2 Procedure

41

Rubber sealing was used to prevent $as leakage from the biotilters and eacli section l i a s
screwed to each other in sis points. Each section of 111c hiotiltrrs Iiad one port

at

ihc top Icvcl

for air sampling and three ports distributed axially for pressure drop iueasureiiiiit:, i 5c.c Fiyiirc:,

3 and 4 . Methanol and a-pinene concentrations in the air were measured ot the inlet. iop 'if
section 2. top of section 3. top of section 4. and outlet of racli biofilter.

Figure 3.2
Photography of the actual experimental set-up. From lefi to right: conrrol panel.
valve control panel and manometer fpr AP measurement. humidification system. biofilters. and
GC

Eflecis of Biomass Growtli ou Pressure Drop iit Biqtiiters

Experimental Procedure

Figure 3.3
Photography of section 1 of the wood cliip biofilter on day 1 1 that s l i o w
axial ports for pressure drop measurement on the right Iiaiid sidr

ilir

gure 3.4
Photography of section 1 of the Nova Inert biofilter on day 1 1 that shows
axial ports for pressure drop measurement on the right hand sidz
-

Effecrs of Biomclss Groirth on Pressure Drop in Biofilrci*.~

fiperimentd Procedure

43

3.3 Packing Materials


In one biofilter, oak and maple hardwood chips (TMP mill, Domtar Packaging, Trenton,
ON) were used as the packing material. The other biofilter was packed with an inert material

(Nova [ne#', NovaBiotec Dr. Fechter GmbH,Berlin, Germany) consisting of porous inflatedglass pellets. The type of wood chips has been successfully used in previous biofiltration
experiments (Mohseni, et al., 1998), and the inert material has been successfully used in
biotrickling filters (NovaBiotec Dr. Fechter GmbH, persona1 communication).

The wood chips and the Nova Inert pellets were sieved through screens of different mesh
diameten. The selected wood chips had a size between 0.7 cm and 3.5 cm, and the Nova Inert
pellets had an average diarneter between 0.5 cm and 1.1 cm. Table 3.1 presents the size
distribution of wood chips and Nova Inert pellets used to pack the biofilters. Small s i x
particles, less than 0.7 cm for the wood chips and less than 0.5 cm for the Nova Inert pellets,
were not used in order to avoid higher pressure drops due to less void space within the bed.
Table 3.1
Size distributions in percentage volume for wood chips and Nova Inert pellets
used to pack the biofilters ( d a stands for not available)

1 Mesh squared pore size /

Percentage (v/v) Wood chips

3.5 > d >2,3 cm

7%

da

2.3 > d > 1.6 cm

33%

nia

1.6 > d > 1.1 cm

40%

da

1.1 > d >0.7 cm

20%

33%

0.7 > d > O S cm

da

66%

/ Percentage (vlv) Nova lnert pellets

--

Each section of the biofilters was packed with 15 L of packing material. Before packing,
macronutrients were supplied to the media with

utr ri cote^

slow-release fertiliser pellets type

270 (Plants Products Co. Ltd., Brampton, ON); 800 g of fertiliser pellets were rnixed with 15 L

of packing matenal for each section. Macronutnent requirements were calculated based on a
CM ratio of 30, recommended for an optimal microbial activity, and based on maximum

loading rates of 150 g methanol/m3 bed h and 40 g ~ - ~ i n e n e / bed


r n ~h for an operating time of
~ffictsof Biomass ~ k t onhPressure Drop in Biojilers

Experimental Procedure

44

120 days with each pollutant (refer to Appendix A for detailed calculations). A micronutrient

solution was added only to the inert material in a ratio of 5 L of solution per 15 L of packing
material. For the case of the wood chips, trace compounds were considered to be supplied by
the packing material itself. Tables 3.2 and 3.3 present the composition and some specifications

of the fertiliser pellets and the composition of the micronutrient solution, respectively. The
addition of both macro- and micronuients was expected to ensure enough biornass growth in
order to register its effects on pressure drop in the biofilters.
Composition and some specifications of the ut ri cote" slow-release fertiliser
Table 3.2
pellets used as a source of macronutrients in the biofilters (Manufatum specifications)

utr ri cote^ pellets type 270 - fertiliser releases 80% of its nitrogen evenly over a 270 day
period at a constant temperature of 25 O C . Faster release at a higher temperature.
Compound

Total Nitrogen

Composition

20 %

10.7%

Ammoniacal nitrogen
Nitrate nitrogen

Available phosphoric acid (P20s)

7%

63%

Soluble potash (K20)


Others not specified

33.1 Preparation and Inoculation


As mentioned before, each section of the biofilters was packed randomly with 15 L of

packhg material. For the case of the Nova Inert biofilter, the size distribution of the packing
materiai mixture for each section was prepared mixing together 5 L of pellets (bulk volume)

with a diameter between 1.1 cm and 0.7 cm, and 10 L of pellets with a diameter between 0.7 cm
and 0.5 cm. The total volume of 15 L of Nova Inert pellets was subsequentiy combined in a
bucket with 5 L of micronutrient solution and 5 L of an inoculum solution. The 15 L of wood
chips for each section was prepared by combining 1 L of wood chips with a mesh diameter
Effects of Biontars Growth on Pressure B o p in Biofilers

k~erimentalProcedure

45

between 3.5 cm and 2.3 cm, 5 L with a mesh diameter < 2.3 cm and N . 6 cm, 6 L with a mesh
diameter < 1.6 cm and > 1.1 cm, and 3 L with a mesh diameter < 1.1 cm and > 0.7 cm. The
wood chips were then mixed in a bucket with 5 L of an inoculurn solution.
Table 3.3

Composition of micronutrient solution added to the Nova Inert packing material


Compound

Concentration (mgL)

Zinc sulphate (ZnS04)

O $44

Ferrous suiphate ( F e S 4 )

2.49

Manganous sulphate (MnSQ)

0.308

Cobaltous chloride (CoC12)

0.404

Sodium chloride (NaCl)

0.254

Magnesium sulphate (MgS04)

2.48

Sodium molibdate (Na2Mo04)

Copper sulphate (CuS04)

Caicium sulphate (CaS04)

1
1

1.26

0.393

0.430

The inoculurn solution was obtained by washing out a wood chip-compost-perlite


medium that had been used in previous methanolla-pinene biofiltration experiments (Mohseni,
1998). The old packing medium was submerged in distilled water, mixed, and let settle

overnight. The supernatant was then decanted and used as the inoculurn solution, which was
added to the new packing matenal, dong with the micronutrient solution (only Nova Inert
pellets). The packing media with the inoculum solution were let to stand in the buckets for 4

days. Just before packing each section of the biofilters with the respective materials, the media

were drained and 800 g of fertiliser pellets were added to each 15 L volume of packing mixture.
Five hundred ml more of the micronutrient solution were added to each section of the Nova
Inert biofilter on day 84.

Effects of Biornass Gmwth on Pressure Drop in Biojilers

Experimental Procedure

46

3.4 Bed Sampling and Pressure Drop (AP)Measuremeats


The amount of biomass developed on the packing materials was assessed in both
biofilters fiom bed composite samples prepared at different bed levels within each section: top,
middle, and bottom. For sampling, each biofilter was dismantled and the bed composite
samples were prepared with samples of pellets or of wood chips taken from 5 different points at
the bed level, as indicated in Figure 3.5. This sampling procedure was perforrned on a periodic
basis, every 4 to 6 weeks. After sampling, the packing material within each section was
remixed in order to ensure a unifonn biomass distribution at the beginning of each undisturbedbed period of operation.
r

Superior
view

Section

Figure 3.5 Sampling points for composite samples at the top,


middle, and bottom of each section of the biofilters

In the wood chip biofilter, new wood chips were added penodically. At the beginning of
the third period (day 76); 2 L of fiesh chips were added to section 1, 1 L to section 2, and 1 L to
section 4. At the beginning of the operation with a-pinene (day 148); 1L of fiesh chips was
added to section 2, 2L to section 3, and 1 L to section 4. This helped to replace the sampled
wood chips and to counterbalance the reduction of bed volume due to sampling.
Pressure drop dong segments of the packed beds were measured as differential pressure
using a digital manorneter (475-1 Mark II, Dwyer Instruments Inc., Michigan City, IN)
connected to a designed valve systern, which permitted selecting of different ports of
Effects of Biomass Growth on Pressure Drop in Biofilrers

Experimentul Procedure

47

measurement dong the biofilters. The manometer minimum detectable differential pressure
was 2.5

Pa (0.01 in water). Before measuring pressure drop, the plastic tubing c o ~ e c t i n gthe

ports to the manometer were cleaned to eliminate condensed water.

3.5 Tracer Studies


Pulse injection tracer tests were conducted using methane (99.0%, Supelco, Inc.,
Bellefonte, PA) as a tracer. The methanol pulse injections were performed manually at the inlet
of the biofilters using a 10 mL pressure-~ok@syringe (VICI Precision Sampling, Inc., Baton
Rouge, LU). The methane sample was injected into the inlet Stream through a septum located in
an axial port of the tubing. The outlet methane concentrations were rneasured every second

using a Total Hydrocarbon Analyser (MSA Baseline 8800 THC analyzer, MSA Canada Inc.,
Downsview, ON). In order to retrieve the measured concentrations in a cornputer system, the
measurements were automatically stored in a datalogger (Ultra-Logger, Lakewood Systems
Ltd., Edmonton, AB), which was connected to the analyser.
The pulse injections were conducted three times in a row, just allowing enough time in
between each injection for the outlet methane concentration to reach the respective baseline of
zero. This ensured the reproducibility of the injections and methane concentration measurement
at the time of the experiments.

3.6 Analytical Techniques


3.6.1 Methanol and a-pinene Concentration Measurements

The concentrations of methanol and a-pinene in the air strearns were measured by gas
chromatography. Axial ports located dong the biofilters were available for automatic air
sampling. The air samples were taken using a vacuum pump (Mode1 400-1901, Barnant
Company, Barrington, IL) and conducted to the GC through 1.6-mm(1116") stainless steel lines.

Sampling and injection were controlled by a 16-position Stream selecting valve (Valco
Instruments Co., Inc., Houston, TX) comected to a 6-position switching valve (Valco
Instnunents Co., Inc.) on the GC. The sampling iines were heated for preventing sanirated gas
condensation, and the sample volumes automatically injected to the GC were 250 PL.

Effects of Biorn<iss Growth on Pressure Drop in BioJ2ter.s

berimental Procedure

48

Methanol and u-pinene concentrations were monitored fiom air samples taken at the
inlet, outlet, and top of the second, third and fourth sections of each biofilter.

A gas

chromatograph mode1 Star 3600 CX (Varian Chrornatography Systems, Walnut Creek, CA) was
used for these analyses. From a total of 4 injection in each port, three were taken to be averaged
and calculate the VOC concentration at the respective port. Helium was used as carrier gas and
the samples were passed through a 15-mmegabore column (0.53 mm ID, J& W Scientific, DB-

9, which was maintained at

120C. The Flame Ionisation Detector was

run with air at 300

mL/min and hydrogen at 30 mLJmin, and the detection temperature remained at 250C.
Appendix B presents the calibration data for measuring methanol and a-pinene in the GC.
3.6.2 Bed Moisture Content and pH

Bed moisture content and pH were measured by duplicate in each section of the
biofilters fiom samples taken after remixing the bed. Bed moisture content was assessed
gravimetrically by the weight lost on heating to 105OC for 24 hours. The weight difference
between the wet bed sample and the dry bed sarnple was considered to be the water content of
the sample. Bed moisture content was then reported as a percentage by weight.
Measurements of pH fiom bed samples were conducted based on Methods for Soi1
Analysis (1996). Approximately 10 g of bed sarnple were placed in 30 mL of nanopure water,
for the case of Nova Inert pellets, and 50 mL of nanopure water for the case of wood chips. The
container was covered with paraffin paper to prevent equilibrium with ambient CO2, and let
stand for 10 minutes. Before introducing the pH probe into the water, the sarnple was stirred
and then the pH measured.

3.6.3 Interparticle Porosity

The interparticle porosity in the beds was measured in each section of the biofilters from
samples taken d e r remixing the bed. In a beaker of known volume and weight, a bed sample
of the sarne beaker's volume was placed in it, weighed, and then filled with water. The volume
of water required to filled out the void space in the beaker was then measured, and the
interparticle porosity was calculated as the quotient of the water volume per unit mass of bed
sample over the bed sarnple volume per unit mass of bed sample.
Eflects of Biomass Growth on Pressure Drop in Biofiters

Erperimental Procedure

49

3.6.4 Biomass Quantification

Biomass quantification was conducted via four different techniques as described below.
The four techniques were applied to two composite samples taken fiom the top, middle, and

bottom of each section of the biofilters. The final biomass quantification was an average of the
two values obtained for each composite sample.
3.6.4.1 Biomass Detachment Technique

The four techniques used involved the detachment of biomass fiom the support medium.
which allowed suspending the biomass in water. The biomass detachment procedure consisted

of the following steps:


1. Weigh empty flask, add bed sample and weigh again.

2. Add 20-25 ml of nanopure water to the bed sample in the flask.


3. Shake for 30 min at 250 rpm (Controlled Environment incubator Shaker. New

Brunswick Scientific Co., Inc., New Brunswick, NJ)


4. Take sample volumes for COD a d o r protein analyses, and decant the remaining

suspension into an aluminium dish.


5. Wash the flask and remaining support matenal with 5 mL of nanopure water and then

pour content into the Al dish.


3.6.4.2 Detached Biolayer

The simplest technique for biomass quantification used in this study consisted of
measuring biomass as g of detached biolayer per g of dry substratum. The mas of detached
biolayer was calculated as the difference in weight between the wet bed sample and the
substratum dned at 105C for 24 hours after detaching the biomass.
3.6.4.3 Total and Volatile Solids (TS & VS)

Total solids and volatile solids were determined in the biomass suspension obtained afier
detaching the biomass fiom the substrata. The definitions and general procedures for measuring
total and volatile solids were followed fiom the Standard Methods for the Examination of Water

and Wastewater (APHA,1992), section 2540 B and section 2540 G, respectively. In general,
the total solids were considered to be the solids remairhg in the sample dish after drying it in an

Effectsof Biomass Growth on Pressure Drop in Biojiters

Experimental Procedure

50

oven at 105OC for 24 hours, and the volatile solids the solids loss on ignition in an oven at
550C. The msults were reported as mg of TS or VS per g of dry substratum.
3.6.4.4 Chemical Oxygen Demand (COD)

COD was determined fiom a 1.5-mL sample of the biomass suspension obtained after
biomass detachment. This sample was diluted with nanopure water to have a total volume of 6

rnL ( 1 5 4 . 5 dilution).

From the diluted 6 rnL sample, 2.5 mL were used for COD

determination.
COD was detemined using a closed reflux, colorimetric technique which was originally
modified fiom the Standard Methods for the Examination of Water and Wastewater (APHA,
1992), section 5220 D. This technique involves the following reagents and procedure (Bagley,

1997):

Reaeents
1. Digestion solution: Add to 500 mL distilled water, 10.2 16 g K2Cr207, primary

standard grade, previously dried at 103C for 2 hours; 167 mL concentrated H2SO4;and
33.3 g HgS04. Dissolve, cool to room temperature, and dilute to 1000 mL.
2. Sulphuric acid reagent: Add 11 g Ag2S04to 2.5 L concentrated sulphuric acid. Let

stand 1 to 2 days to dissolve ApS04.


3. Potassium hydrogen phthalate (KHP) standard: Lightly cmsh and then dry KHP to

constant weight at 120C. Dissolve 680 mg in distilled water and dilute to 1000 mL.
This solution has a theoretical COD of 800 mg/L.

I?mduts
1. Place 2.5 of diluted sample into a 16x 100 mm COD digestion tube. Add 1.5 mL of

digestion solution and 3.5 mL of sulphuric acid reagent. Tightly cap the tubes, mix well,
digest for two hours at 150C in a block heater, and let them cool.
2. Prepare a blank for every group of samples, adding 2.5 ml of distilled water instead

of sample. Al1 the digestion steps are identical.


3. Prepare a calibration curve using the KHP standard solution. Prepare standards of 20,

40, 80, 200, 400,600, and 800 mg CODL Use 2.5 mL of standard solution instead of
sample and follow al digestion steps. A blank should be prepared as well.
Effecfsof Biomuss Growth on Pressure Drop in Biojlters

Experimentuf Procedure
4.

51

After the tubes have been cooled, measure the absorbance at 600 nrn with a

spectrophotometer (Spectronic 20, Bausch & Lomb, USA).


5. Prepare an 8-point calibration cuve by plotting the COD concentration versus the

measured absorbance, including the blank.


6. Detemine the sample COD concentration fiom the calibration curve, reading up from

absorbance to the curve and across to COD concentration.


7. The sample COD concentration (mg CODL) was further divided by the dilution

factor of 0.25, and corrected for the total volume of biomass suspension fiom which the
sample was onginally taken. The final result was reported as mg COD per g of dry
substratum.
3.6.4.5 Total Protein

The protein content in sarnples from the biomass suspension was measured using a
colorirnetric method (Lowry et al., 195l), based on the Folin reaction (Famgia, 1999). The
general procedure is presented below.
1. One mL of sample was added to 5 ml of reagent C (20 g Na2C03 in 1 L of 0.1 N

NaOH, mixed with 0.5 g CuS04.5H20 dissolved in a 1% (w/v) aqueous solution of

sodium tartrate) in a ~ a c h @
culture tube and mixed well.
2. The mixture was allowed to stand for 10 minutes at room temperature.

3. Another 0.5 mL, now of Folin reagent (diluted Folin and Ciocalteu's phenol reagent,
10 mL to 90 mL of deionized distilled water) was then added to the mixture and mixed

immediately.
4.

The absorbance (Atow) of the samples was measured at 750 nrn with a

spcctrophotometer. Bovine serurn albumin was used as the standard solution.

5. Protein content measurements were corrected for humic substances present in the
biomass suspension.
Humic substances were determined with the following steps (Famgia, 1999):
1. One mL of sample was aded to a borosilicate g l w 10 mL test tube. A blank was

prepared with 1 mL distilled water.

E-ects of Biomass Growth on Pressure Drop in BiofiZters

Experimentul Procedure

52

2. Five mL of alkaline buffer solution (20 g Na2C03dissolved in 1 L of 0.1 N NaOH)


were added to the test tubes and mixed on vortex.

3. The mixture was allowed to stand for 10 minutes at room temperature.


4. 0.5 mi,of Folin reagent were added and mixed imrnediately on vortex.
5 . The mixture was allowed to stand for at least 30 minutes more at room temperature.

6. Absorbante (Ablind)was measured at 735 nm in a spectrophotometer (SP6-500UV


Spectrophotometer, Pye Unican, England). A standard humic acid solution was used for
the calibration curve.

7. Finally , the calculation of absorbance due to proteins and humic substances was
conducted as follows:
Aprokin=
Ahumis

1 WAtotai - Abiind)

= Ablind - 0-2Aprotsin

M e r converting these values to mg of proteins, the final results were given as mg of


protein per g of dry substratum.

The protein and humic acid analyses were conducted with 1 or 2 mL of original biomass
suspension, which were diluted differently for each sampling day in accordance to the biomass
concentration. From the diluted sample, 1 mL was taken for the protein and humic acid tests,
respectively. The IrnL or 2 mL of biomass suspension were taken fiom a 3 mL sample
confonned by mixing 1.5 mL fiom the biomass suspension of a first composite bed sample, and
1.5 mL fiom the biomass suspension of a second composite bed sample.

- -

E w s of Biomuss Growth on Pressure Drop in Biofiters

Results

4 RESULTS
Methanol was used as the pollutant during most of the operation period of both
biofilters. The results for the first 150 days of operation using methanol are presented in
subsection 4.1. Results from the subsequent biofiltration of a-pinene for 6 1 days are reported in
subsection 4.2.
4.1 Biofiltration of Methanol
4.1.1 Performance
4.1.1.1 Overall Methanol Removal

During the first 150 days, both biofilters were operated at similar methanol loading rates
and methanol inlet concentrations as showri in Figures 4.1 and 4.2. Figure 4.1 presents the
performance of the biofilters in terrns of inlet and outlet concentration versus tirne, and in Figure
4.2 methanol loading rate and percentage removal are plotted over time for both biofilten.

In the first 20 days of operation, the methanol loading rate was gradually increased from
15 to 100 g methanol/m3 bedh in order to allow an active microbial community to grow.
Methanol inlet concentrations and empty bed retention times (EBRT)varied accordingly up to
900 ppmv and 42 s, respectively. Values for the first 5 days of operation were not registered
because of operational problems with the OC. From day 20 to day 40, an attempt was made to
maintain a constant loading rate around 100 g merhanoV m3 bedh since methanol removals
higher than 98 % were observed and this loading rate was considered high enough to provide
extensive microbial growth. After day 40, the loading rate was gradually increased, and around
day 70 the system was operating at loading rates as high as 250 g methanol/m3 bedh. Before
this increase, high variations in loading rates and methanol inlet concentrations were registered
because of line plugging due to a-pinene accumulation on the wall of the lines from previous

biofiltration experirnents. When loading rates of 200 g methanol/m3 bedm were reached, the
biofilters were running at its maximum capacity, and wide loading variations were also
registered.

M e r day 80, the loading rate was kept relatively constant around 120 g

methanol/rn3bed/h, but also decreases were observed due to plugging in the delivery lines.

E w s of Biomass Growth on Pressure Drop in Biofilers

Results

54

20

40

60

1O0

80

120

160

140

Time (days)

Figure 4.1
Inlet and outlet concentrations of the Nova Inert and wood chip biofilter during
the biofiltration of methanol at 40C bed temperature

20

40

60

80

t00

120

140

160

Tirne (days)

Figure4.2 Nova Inert and wood chip biofilters performance durhg the biofiltration of
methanol at 40 O C bed temperature

Results

55

Figures 4.1 and 4.2 also show that the two biofilters performed similarly. AAer the first
40 days of operation, methanol inlet concentrations varied in the range of 600 to 1200 ppmv,
and loading rates varied from 100 to 250 g methanol/m3 bedh; however the removal efficiency
remained higher than 95 % at EBRT of 27 to 30 s in both biofilters.
4.1J.2 Axial Profiles of Metbanol Removal
As presented in the axial profiles of Figures 4.3 and 4.4, in both the Nova Inert and the
wood chip biofilter, most of the methanol was removed in the first 3 sections (around 0.7 to 0.8

m bed depth) of the biofilters during the 150 days of experiments. Each of these profiles was
taken at the end of an undisturbed-bed operation period. As s h o w , at the end of the first
undisturbed-bed period (Nova Inert: day 55; wood chip: day 49), the first two sections removed
up to 95 % of the incoming methanol in both biofilten. But in the second undisturbed-bed
period, as the rnethanol loading rates increased, methanol removal in section 3 began to
increase.

Figure 4.3
Typical methanol axial concentrations dong the Nova Inert biofilter before
sampling and remixing the bed during the biofiltration of methanol. A dashed line indicates the
approximate location of each section.

Effecfsof Biomuss Growth on Pressure Drop in Biofiters

Figure 4.4
Typical methanol axial concentrations dong the wood chip biofilter before
sampling and remixing the bed during the biofiltration of methanol. A dashed line indicates the
approximate location of each section.
The profiles shown in Figure 4.3 and particularly in Figure 4.4 reflect variations in
methanol loading rates in the system. As shown in Figure 4.4 the variation was great for day
104 the inlet methanol concentration in the wood chip biofilter was lower than for the other days

because of a decrease in loading rate experienced at this tirne. A decrease in removal capacity
in section 1 of the wood chip biofilter occurred at the end of the last three undisturbed-bed
operation periods (days 76, 104, and 139) (see Figure 4.4). This removal decrease coincided

with both less biomass growih and lower pressure drop in this section over these last 3 periods
of operation.
4.1.2 Biomass Accumulation in the Packed Beds

Biomass concentration in the packing material was measured penodically by sampling


the bed of each section at the top, middle, and bottom. During the methanol feeding penod, the
bed of each biofilter was sampled four times, once at the end of each undisturbed-bed operation
Effects of Biomass Growth on Pressure Drop in Biojlters

period. The wood chip biofilter was sampled on days 49,76, 105, and 148, while the Nova lnert
biofilter was sarnpled on days 56,84, 111, and 148. The first 3 sarnpling days for both biofilters
were different to facilitate handling and testing of samples. Detached biomass was measured
using four techniques. A water suspension of detached biomass was prepared fiom which
biomass was quantified as Chemicai Oxygen Demand (COD),Total Solids (TS), and Total
Volatile Solids (TVS). In addition, by difference in weights between wet bed sample and clean
dry packing material, biomass was quantified as g of biolayer, which included biomass and
water present in the bed sample. Al1 biomass measurements were norrnalised to g of dry
substratum where dry substratum denotes dry packing material, either Nova Inert pellets or
wood chips.

4.1.2.1 Biomass Quantification as Detached Biolayer


When quantifying biomass as g of biolayed g dry substratum, the term biolayer

considers the wet rnicrobial mass detached from the substratum together with the water content
of the packing material. The profiles of detached biolayer are s h o w in Figure 4.5 for the Nova
Inert biofilter and in Figure 4.6 for the wood chip biofilter. These biomass profiles are in
accordance with visual observations of biomass development on the beds, and pressure drop
measurements, and therefore, were considered suitable for descnbing biomass accumulation in
the biofilters.
Figure 4.5 clearly shows that most of the biomass accumulated in the first half of the
Nova Inert bed. The decreasing biomass concentration along the bed are in agreement with
observations of biomass development on the packing material. In the wood chip biofilter these
concentration profiles tended to have the same decreasing trend dong the bed; however, there
are two concentration peaks on day 76, which account for the greatest accumulation of biomass
at the top of sections f and 2 of the wood chip biofilter. Moreover, a significant difference
between the amount of biomass accumulated at the top and middle and bottom of each section
can be obsewed in the profiles of both biofilters, but it is more evident in the wood chip biofilter
(Figure 4.6). The biomass concentrations on the wood chip bed are roughly 3-4 fold higher than
Effects of Biomass Growth on Pressure Drop in Biojlters

in the Nova Inert bed; however, in both biofilters the biomass profiles show uneven
accumulation along the bed and in between sections. High concentrations at the top of section 1
in the Nova Inert biofilter (Figure 4.9, and at the top of section 1 and 2 on day 76 in the wood
chip biofilter (Figure 4.6 below) correlate to the observed formation of thick biomass layers on
the top of these sections.

I
I
I
I
I

I
I
I
I

I
I
I
t
I
I
I
I

-+Day SS
+Day 84

-+ Day I l 1

+Oay 148

I
I

section 4

Figure 4.5
Biomass concentration profiles as detached biolayer along the Nova Inert biofilter in 4
different days during the biofiltration of methanol. A dashed line indicates the approximate location of
each section.
A relative increase of biomass over time in al1 the sections of both biofilten is s h o w by

the detached biolayer measurements, which is consistent with the continuous removal of the
pollutant along the beds.

Effects of Biomass Growth on Pressure Drop in Biojiters

Results

59

+Day 49

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

+Day

76
Oay 105

*Day

148

0.9

Bed depth (m)

Figure 4.6
Biornass concentration profiles as detached biolayer along the wood chip biofilter in 4
different days during the biofiltration of methanol
4.1.2.2 Biomass Quantification as Chernical Oxygen Demand (COD), Total Solids (TS),

and Total Volatile Solids (TVS)


Biomass suspensions from bed samples collected at different bed levels were analysed
for COD,TS, and TVS to evaluate these techniques for measuring biomass concentration on the
packing materials compared to biomass quantification as g of biolayedg dry substratum. These

profiles for both biofilters are presented in Appendix C.


The COD,TS, and TVS profiles in both biofilters show an increase of biomass with
tirne, which correlated well to the other biomass quantification technique results. However, for
the last two sampling days, the COD and TS/TVS measurements indicated more biomass in

section 3 and 4 than expected and than visually assessed, which contradicted the other results of
biomass quantification and pressure drop. This might be due to interference fiom the wood
chips in these techniques. Higher values could be explained by the presence of substances and
fibres released from the wood chips during the preparation of the biomass suspension since
wood chips were being degraded and were loosing their firm structure. Higher biomass

concentrations as solids were measured in the wood chip bed, 1.5- to 44old higher, than in the
Nova Iwrt bed. The solids concentration in the biomass suspension from the wood chip
Effects of Biomass Growth on Pressure Drop in BioBIters

Results

60

samples was expected to be higher since there was visibly greater biomass accumulated on the
wood chips than on the Nova Inert pellets.
Both the TS and the TVS profiles present a similar trend in both biofilters. For the Nova
Inert biofilter the average TVSITS ratio for the analysed sarnples was 0.582 (std. deviation =
0.096), indicating that more than half of the TS were organic solids. This ratio was expected to
be closer to one since most of the biomass material is largely organic. For the wood chip
biofilter, the average ratio TVS/TS of the analysed sarnples was 0.775 (std. deviation = 0.185),
higher probably due to organic materials detached fiom the wood chips.

The COD,TS, and TVS profiles show some of the high concentrations at the top levels
of the sections of the biofilters, and thus, the uneven biomass accumulation, in a similar way as
was represented by the detached biomass measurements. However, the highest biomass
concentrations corresponding to the thick biolayer formed at the top level of section 1 were not
measured neither by the COD nor by TS/TVS tests. This is more clearly illustrated by the
comparative plot of the results in Figure 4.7 and 4.8, which show the results of COD and TS
measurements versus g of biolayer per g of dry substratum for the Nova Inert biofilter. The
majority of the data points tend to describe a proportional relation between the measurements,
especially for the case of TS (Figure 4.8). Nevertheless, there are some data points (marked

within a rectangle) that are the highest biomass concentrations measured as g of biolayer per g
of dry substratum that do not follow this trend. These data points correspond to the peak
biomass concentrations observed and measured as detached biomass at the top level of the first
and second section of the Nova Inert biofilter. Since these high values of biomass concentration
as detached biomass have similar COD and TS values than lower values of detached biomass
concentration, this indicates that the COD and TS,and correspondingly, the TVS measurements
fail to quanti@ this type of biomass accumulated at the top of the bed, which is the area where
biomass had a great impact on pressure drop in the biofilter. Similar results were obtained in
the wood chip biofilter (see Appendix D).

EBcts of Biomass Growth on Pressure Drop in Bioflters

II
X

0.5

X X X

2
Q biolayerlgdry rubtratum
1.5

2.5

Figure 4.7
Comparison between biomass measurements as mg COD/$dry substratum and as g
biolayerlg dry substratum in the Nova Inert biofilter during the biofiltration of methanol

Figure 4.8
Comparison between biomass measurements as mg TSIg dry substratum and as g
biolayerfg dry substratum in the Nova Inert biofilter during the biofiltration of methanol

Effects of Biomass Growth on Pressure Drop in Biojifers

3.5

Results

62

4.1.2.3 Patterns of Biomass Accumulation

Visual assessrnent of the beds during sampling indicated that biomass accumulated in an
uneven fashion. On the first two occasions when the beds where sampled (Nova Inert biofilter:
day 56 and 84; wood chip: day 49 and 76), a layer of biofilm, approximately 1 cm thick, was
observed at the top level of the first and second sections of both biofilten.

Biomass

accumulation was greater on the wood chips than in the Nova Inert pellets, but in both biofilters
biomass accumulated in a patchy way, especially at the top of the first sections and near the
walls of the biofilters. Inside the beds, biomass also tended to grow near the walls leaving
central areas dry, particularly during the first 100 days of operation. During the last 50 days of
operation, previous mixing of the beds might have hefped to obtain a more uniform biomass
accumulation in the central inner parts of the beds. The uneven biomass growth in the beds is
illustrated in Figures 4.9 and 4.10, where the top levels of the four sections of the biofilters are
shown for day 84 for the Nova Inen biofilter and for day 74 for the wood chip biofilter.

the
end of the second undisturbed-bed period of operation (day 84)

Eficts of Biomass Growth on Pressure h o p in Biofilers

Results

63

the

end of the second undisturbed-bed period of operation (day 74)

4.1.3 Pressure Drop (AP)

Pressure drops, both total and at different bed levels, dong the Nova Inert and wood chip
biofilters were regularly monitored. During the operation of the biofilters, in between two days
of bed sampling and mixing, the bed was not disturbed. Consequently, 4 undisturbed-bed
periods of operation were allowed since the beds were sampled 4 times in total.
4.1.3.1 Total Pressure Drop

As shown in Figure 4. 11, the total pressure drop dong both biofilters began to increase

steadily &er 30 days of operation. Since the beginning, greater pressure drops were measured

in the wood chip biofilter, and pressure drop increased more rapidly in the wood chip biofilter
than in the Nova Inert biofilter. The highest pressure drop values were registered during the
second undisturbed-bed period of operation, when the formation of a biomass layer at the top
levels of sections 1 and 2 of both biofilten caused pressure drops as high as 2600 Pidm bed in
the wood chip biofilter, and 550 Pdm bed in the Nova Inert biofilter. In the case of the Nova
.

.
.

-..--.

Effects ofBiomms Growth on Pressure Drop in BioJfters

Results

64

Inert biofilter, a steady increase in pressure &op over time was observed after remixing the bed
at the end of each period. However, in the wood chip biofilter the steady build-up in pressure
drop was only observed in the first two undisturbed-bed periods. In the last two penods of
operation, &er observing a steady increase in pressure drop, the pressure drop remained
relatively constant and/or with small increases. This pressure drop behaviour in the wood chip
could be related to the way biomass accumulated, and to the changes that the wood chips were
experiencing because of self composting and supplying nutrients for biomass growth.
Pressure drop was on average 6-fold higher in the wood chip biofilter than in the Nova
Inert biofilter, consistent with more biomass growth on the wood chips and compaction of the
bed. Figure 4.1 1 also shows that the pressure drop peaks before mixing were different. The
total pressure drop, however, decreased almost to zero afier remixing, indicating that an even
biomass distribution within the bed was a controlling factor in pressure &op.
-

Period 4

+Fkvalnert biofilter
+Wood chip biediter

Figure 4.1 1 Total pressure &op dong the Nova Inert and the wood chip biofilters during the
biofiltration of methanol. Bed remixing points are indicated by continuous lines for the Nova
Inert bed and by dotted lines for the wood chip bed.

Effectsof Biomuss Growth on Pressure Drop in Biofilers

4.1.3.2 Pressure Drop Axial Profiies

In order to assess how pressure &op was being afTected dong the bed of the biofilters,
pressure drop profiles were measured on a daily basis. Cumulative pressure drop profiles at the
beginnhg and at the end of the 4 undisturbed-bed periods are presented in Figures 4.12 and 4.13
for the Nova Inert and the wood chip biofilter, respectively. In general, for both biofilters,
similar profiles were observed, where most of the pressure drop (around 90 %) was caused in
the first half of the bed, i.e. sections 1 and 2, during al1 the 4 undisturbed-bed periods.
Figures 4.12 and 4.13 also show higher pressure drops in the wood chip biofilter in
cornparison to the Nova Inert biofilter. The effects that mixing had on the decrease of pressure
drop is also evident in these Figures. M e r remixing, the pressure drop was significantly

reduced to a value similar to the initial value of the period. AAer mixing, most of the pressure
drop registered at the beginning of each period was caused in the first 30 cm of bed (section 1)

of both biofilters, where more biomass had accumulated.

Figure 4.12 Cumulative pressure &op profiles dong the Nova Inert biofilter at the b e g i ~ i n gand
at the end of each undishirbed-bed period of operation. A dashed line indicates the approximate
location of each section

Effects of ~iom&s~ r o w t hon Pressure Drop in Bioflters

Results

66

+-&y
-A-

19 (period 1)

D ~50
Y (pend 2)

+Day 77 (period 3)
+-Day
106 (pend4)
-

. -.

Figure 4.13 Cumulative pressure drop profiles dong the wood chip biofilter at the beginning
and at the end of each undisturbed-bed period of operation. A dashed line indicates the
approximate location of each section
Figure 4.13 shows that in the wood chip biofilter at the end of penod 4 (day 146) and in
the first 30 cm of bed (section 1) pressure drop did not increase as significantly as in the
previous periods. This was consistent with less biomass growth and less compaction in this
section of the wood chip biofilter. This could be an indication that the microbial activity was
reduced in this section.
4.1.3.3 Pressure Drop vs. Aidlow Rate

As a way of evaluating flow charactenstics through the biofilters, pressure drops at


different airflow rates were measured on different days during the undisturbed-bed operation
periods. Figure 4.14 shows the total pressure drop dong the Nova Inert biofilter vs. airfow rate

during the third undisturbed-bed operation period. Similar curves for the same period of
operation are show in Figure 4. 15 for the wood chip biofilter.

Effects of Biomass Growfhon Pressure Drop in Biofikers

Results

20

40

60

80

100

120

140

Air flow rate (Umin)

Figure 4.14 Total pressure drop in the Nova Inert biofilter, as a function of airflow rate.
measured on different days during the third undisturbed-bed period of operation
The semiparabolic curves of total pressure drop vs. airfiow rate were obtained during the
entire period of operation of the biofilters with methanol. As s h o w in Figure 4.14 for the Nova
Inert biofilter during period 3, the curvature of these lines tended to increase accordingly to
higher total pressure drops as time went on and biomass accumulation took place. Similar
results were obtained for al1 the undisturbed-bed operation periods, and also in the wood chip
biofilter. These results indicate that increased biomass accumulation had an effect on the
pressure drop in the biofilters. It is probable that biomass increased the kinetic energy losses
dong the bed, causing an increase in the curvature of these lines at higher flow rates, and higher
pressure drops as biomass accumulated on the packing materials. Compaction may have also
affected pressure &op in the woodchip biofilter.

Eflects of Biomuss Growfhon Pressure Drop in Biofiliers

Results

68

Figure 4.15 Total pressure drop in the wood chip biofilter, as a fnction of airflow rate,
measured on different days during the third undisturbed-bed period of operation
4.1.3.4 Bed Compaction

The wood chip bed, in contrast to the Nova Inert bed, underwent compaction, which was
calculated as the difference between initial and final bed height over initial bed height for each
period. Even though compaction took place during the operation with methanol. the greatest
compaction, around 7 % of the totd height of the bed, occurred in the last undisturbed-bed
operation penod (day: 106-148), possibly due to the degradation of the wood chips. As shown

in Figure 4.16, compaction of the wood chip bed during the other penods was around 3.5 %.
Most of the compaction took place in the first and second sections of the biofilter, except in the

fust undisturbed-bed period when sections 3 and 4 also compacted. As showed by Figure 4.16,
compaction of each section did not foliow any specific trend. Bed remixing and different
biomass accumulation in each section could have cuased different compaction patterns in each
section of the biofilter.

Effectsof Biornass Growth on Pressure Drop in Biofikers

Results

69

lnitid bed
depth:1.O3m

Initial bed
depth: 0.97 m

Figure 4.16 Percentage bed compaction measured as percentage of initial bed height in the
wood chip biofilter during the four undisturbed-bed operation penods. The duration of each
period appears in parenthesis on the x axis.

4.1.4 Bed Moisture Content and Interparticle Porosity

Two parameters that are related to biomass accumulation and pressure drop are bed
moisture content and bed porosity. Moisture content was measured in each section of the
biofilters fiom mixed samples taken fiom each section at the end of each undisturbed-bed
period. Figures 4.17 and 4.18 show the moisture content at the end of each period in the Nova
Inert and wood chip biofilters, respectively. As suggested by the tests of biomass quantification

as detached biolayer, which also included the water retained in the bed, the moisture content in

each section of the beds tended to increase throughout the experiment. A relatively constant
moisture content was measured in the e s t section of the wood chip biofilter, where relatively
low biomass concentrations were measured during the last two periods. Higher moisture
content in the wood chips than in the Nova Inert material was probably due to the presence of
more biomass in the wood chip biofilter, and to higher water retention by the wood chips.

Effects of Biomass Growth on Pressure Drop in Biojiters

Results

70

Section 4

Figure 4.17 Moisture content in each section of the Nova Inert biofilter at the end of each
undisturbed-bed period of operation. Error bars are standard deviations.

Figure 4.18 Moistwe content in each section of the wood chip biofilter at the end of each
undisturbed-bedperiod of operation. Error bars are standard deviations.

Effkcts of Biomuss Growth on Pressure Drop in ~iofilteri

As biomass grows on the bed packing material, the interparticle void spaces are
occupied by biomass and, consequently, the porosity of the bed is diminished. As seen fiom
Figure 4.19 for the Nova Inert bed, the interparticle porosity tended to decrease over time as
more biomass accumulated. Figure 4.20 also indicates a similar effect of biomass growth on
porosity in the wood chip biofilter; however, the high standard deviations in the wood chip bed
do not allow a final conclusion. The high standard deviations in the measurements of porosity
in the wood chip biofilter might have been due to the less unifom shape and size of the chips
which interfered with the technique used to measure porosity since the bed structure could not

be maintained.

Section1

Section2

Section 3

Section 4

Figure 4.19 Interparticle porosity in each section of the Nova Inert biofilter at the end of each
undisturbed-bed period of operation (biomass-fiee porosity = 0.52). Error bars are standard
deviations.

Effectsof Biomuss Growth on Pressure Drop in Biofiters

Results

72

Figure 4.20 Interparticle porosity in each section of the wood chip biofilter at the end of each
undisturbed-bed period of operation (biomass-fiee porosity = 0.62). Error bars are standard
deviations.
4.2 Switchover from Metbanol to a-pinene
4.2.1 Switchover Period

The pollutant shift fiom methano1 to a-pinene was made after sampling and remixing
the bed, and &er allowng the biofilters to reach a pseudo-steady state removing methanol.
Once this state was reached, the system was fed with a-pinene. AAer bed sarnpling and
remixing, under an average methanol loading rate of 123 - 132 g methanoll m3 bedh and an

EBRT of 30 s, the steady operation was reached in 22.5 hours in the Nova Inert biofilter, and in
19.5 hours in the wood chip biofilter (see Appendix E). Under steady operation, 96-98 %
methanol removal was achieved in both biofilters. Under transient conditions, in the first 2 to 3

hours after the start-up, methanol removals around 80- 95 % were observed, which was
attributed to absorption of methanol into water. About 6 to 7 houn after the s*t-up, methanol

removals dropped off, but they began to increase again with tirne, indicating that biological
activity was taking place.
Effects of Bionass Growth on Pressure Drop in Biojiters

After the switchover to a-pinene, the acclimation period for the degradation of a-pinene
until a pseudo-steady state was reached, took about 64 to 65 hours in both biofilters. During this
period, the biofilters were d n g at an a-pinene idet concentration of 22 ppmv and EBRT's of
68-70 s, which corresponded to an a-pinene loading rate of 6-7 g a-pinene/m3bed/h. An a-

pinene removal of 100 % was achieved. The results of this period are presented in Appendix F.

4.2.2 Ovenll Performance, Pressure Drop, and Biomass Growth

AAer the acclimation period of 65 h and until day 166, both biofilters operated with apinene removals above 90 %, a-pinene inlet concentrations of 16-21 ppmv and EBRT's of 34-

42 s, that represented a loading rate of 7- 1 1 g ~ - ~ i n e n e /bedm


m ~ (see Figure 4.21). On day 1 55
the recirculating water fiom the Nova Inert biofilter was accidentally overheated, and the bed
temperature remained at 55C for approximately 5 h, which caused the microbial activity to
diminish almost completely; however, the microorganisms recovered and achieved a 100 %
removal capacity within 4 days. A decrease in removal down to 40 % around day 167 was due
to a 21-hour shut down of a-pinene while still running humidified air. AAer this penod, the

wood chip biofilter did not reach removal efficiencies higher than 90 % and its performance
decreased. In contrast, the Nova Inert biofilter continued to perform satisfactorily, recovering
after bed sampling and remixing. Nevertheless, at the end of the experiments the removal
effciency declined as in the case of the wood chip biofilter.
During the biofiltration of a-pinene, the a-pinene inlet loading was maintained relatively
constant at 10 g a-pinenelm3 bed/h, except at the end of the operation penod. Line clogging
problems and removal efficiencies lower than 90 % in the wood chip biofilter prevented an
increase in the loading rate. This period was charactensed by unstable removals, and a slow

decrease in performance, which was most evident in the wood chip biofilter. This behaviour
seemed to be related to the process of wood chip degradation. A change in the wood chip
appearance was observed during the biofiltration of a-pinene; the wood chips darkened at the
surface, and in both biofilters, white fimgal materid formed.

Efjcts of Biomass Growth on Pressure Drop in Biojilrrs

Results

74

Figure 4.2 1 Performance of the Nova Inert and wood chip biofilters removing a-pinene, after
150 days of operation with methanol
Figure 4.22 shows that the total pressure drop in both biofilters remained relatively
constant, except for an increase observed during days 150 to 160, which coincided with the
highest removal eficiencies of both biofilten during the biofiltration of a-pinene. Pressure
drop in the wood chip biofilter was lower than in the Nova Inert biofilter, even after sampling
and remixing the beds on day 182, contrary to what happened during the biofiltration of
methanol. M e r remixing, pressure drop decreased in the wood chip biofilter, but in the Nova
Inert biofilter it rernained as high as before. This suggests that biomass in the Nova Inert
biofilter might have been uniformly distributed after 5 previous remixings, and that it might
have reached such a concentration where remixing did not have any impact on pressure drop.
This observation is supported by biomass quantification presented as detached biolayer in

Figure 4.23 for both biofilters. As seen fiom this Figure, biomass dong the Nova Inert biofilter
remained relatively constant during this 60-day penod and more evenly distributed dong the
biofilter, although in the last half of the bed slightly lower concentrations were measured. The
biomass concentrations in the Nova Inert bed measured during the biofiltration of a-pinene
ranged fiom 2.5 to 1.5 g biolayerl g dry substratum, which are close to the range measured
Effects of Biomass Growth on Pressure Drop in Biofilers

Results

75

before switching over fiom methanol to a-pinene (2.9-0.8 g biolayed g dry substratum).
Pressure drop profiles along the Nova Inert biofilter (see Appendix G) were similar to those
observed during the biofiltration of methanol, with most of the pressure drop occurring in the
first half of the bed.
Pressure cirops registered in the wood chip biofilter were lower compared to previous
values measured during the biofiltration of methanol, and compared to values of the Nova Inert
biofilter. This dong with the decreased performance of the wood chip biofilter during the
biofiltration of a-pinene suggests that the microbial activity in the wood chip bed was
diminishing. Biomass, quantified as detached biolayer and presented in Figure 4.23, remained
at an average value of 6-7 g biolayerl g dry substratum as at the end of the methanol biofiltration

period. Pressure drop profiles of the wood chip biofilter, presented in Figure 4.24, show that

during the last undisturbed-bed period (period 6), the pressure drop did not increase as rapidly
along the bed depth as it did during penod 5. This agrees with the biomass concentration
profiles along the wood chip bed presented in Figure 4.23, and which biomass concentration
increases slowly along the bed.

Figure 4.22 Total pressure drop in the Nova Ineri and wood chip biofilters during the
biofiltration of a-pinene
Effects of Biomass Growth on Pressure Lkop in Biofllters

Results

76

Figure 4.23 Profiles of biornass quantification as detached biolayer dong the Nova Inert biofilter
and the wood chip biofilter at 2 bed sampling days during the biofiltration of a-pinene

Figure 4.24 Cumulative pressure drop profiles dong the wood chip biofilter at the beginning
and at the end of each undisturbed-bed operation period during the biofiltration of a-pinene
Effects of Biotnass Growth on Pressure Drop in Biofilers

Results

77

Profiles of biomass quantified as COD, TS, TVS, and total protein content are shown in
Appendix H.

4.3 Tncer Experiments


In order to demonstrate whether air channelling was taking place, and to measure the bed
porosity, tracer tests were conducted across the total bed of both biofilters. Methane pulse
injection tests were perfomed on day 126 during the biofiltration of methanol; before and afier
remixing the bed for switching over to a-pinene, on days 147 and 148; and on day 167 during

the biofiltration of a-pinene. The methane pulse injections were carried out in triplicate in both
biofilters. Unfomuuitely, the response curves obtained from dl the tracer tests seem to imply the

opposite of what was expected based on other results and on the physical characteristics of the
beds. Further results and discussions are presented in Appendix K.

Effects of Biomms Growth on Pressure Drop in Biofiters

Discussion

5 DISCUSSION
5.1 Evaluation of Biomass Quantification Techniques

Quantification of biomass accumulation on the packing materials involved both


detaching biomass from the support medium and measuring the amount of microbial m a s .
Quantification as detached biomass was attempted by measuring Chemical Oxygen Demand
(COD), and Total Solids and Total Volatile Solids (TS and TVS) fiom the detached biomass

suspension. By difference in weights between the bed sample and the dry bed sample after
biomass detachment, microbial mass was also quantified as detached biolayer, which inciuded
the bed moisture content.
These four techniques were selected based on literature since they follow a relatively
simple procedure and they do not require sophisticated equipment. Moreover, it was thought
that these techniques could measure indirectly, for the case of COD, TS and TVS, and directly,
for the case of detached biomass, the total amount of biomass growing on the packing material

responsible for airfiow energy losses. Since the goal of quantifiying biomass was to relate this
to the volume occupied by the ce11 m a s , techniques that measure active biomass, such as ce11
counting or measurement of adenosine triphosphate (ATP), would have not been useful since
the activity of the biomass excludes biomass constituents such as EPS and dead microbial mass.

The same applies to biomass techniques based on specific biofilm constituents like proteins or
polysaccharides whose composition may vas, on different packing materials and within the
biofilm and EPS (Cooksey, 1992; Famgia, 1999), which hence may add more inaccuracy to the
rneasurements.
Composite sampling of the bed at the top, middle, and bottom of each section
satisfactorily described biomass accumulation in terms of both quantity and distribution on the
packing material, based on visual assessrnent and pressure drop measurements. Biornass
quantification by detached biomass (g biolayedg dry substratum) offered the best description of
the differences in the amount of biomass accumulated on the media and of its uneven
distribution within the beds. Moreover, biomass accumulation patterns observed in the beds and

the pressure drop profiles dong the beds (Figures 4.12 and 4.13) correlated with the detached
Effects of Biomass Growth on Pressure Drop in Bioflters

Discussion

79

biomass axial profiles (Figures 4.5 and 4.6): higher pressure drops were measured in the first
hdf of the beds where higher biomass concentrations were also quantified.
The COD tests, and especially the TS and TVS tests, did not correlate with visual
assessrnents of biomass growth in the beds and with pressure drop axial profiles, particularly, at
the highest concentrations of biomass observed at the top levels of the first two top sections of
both biofilters during the first two bed-undisturbed periods of operation.

Low biomass

concentrations were measured as COD, TS, and TVS at these top leveis while high
concentrations of biomass were visually detected and quantified as detached biomass (see
Figures 4.7 and 4.8). This thick biolayer was not measured as COD, TS or TVS probably
because of the high water content of the EPS,which were suspected to be its major components;
however, this biolayer increased the pressure drop significantly by decreasing the bed void
volume. Biomass water content in this thick layer was as high as 27-59 g watedg TS in the
Nova Inert biofilter (compared to 11-1 3 g water/g TS in the rest of the bed), and as high as 97-

122 g water/g TS in the wood chip biofilter (compared to 20-22 g water/g TS in the rest of the
bed), which supports these results.
Aside fiom biomass underestimation by COD, TS and TVS at high water content on the
top levels, these techniques showed results in agreement with the detached biomass profiles
dong the rest of the beds. Therefore, the COD, TS and TVS techniques may be useful for
biomass measurement in biofilters with low water content biofilm, probably induced by medium
to low pollutant loading rates. The COD rneasurements, and in a lesser extent those of solids,
seem to depend on the type of packing material. Interference fiom the wood chips might have
increased the COD and solid concentrations, especially during the last two operation periods
during the biofiltration of methanol (Appendix C), when the wood chips were degrading. The
measurements of biomass as COD,TS, and TVS are in agreement with each other as shown by

the proportionality relation in Figures 5.1 and 5.2, which compare the COD and TSlTVS
measurements in the Nova Inert and in the wood chip biofilters.

Effects of Biomass Growth on Pressure Drop in Biofiters

Discussion

80

Comparison between biomass measurements as mg CODIg dry substratum and as


mg TS (TVS)/gdry substratum in the Nova Inert biofilter during the biofiltration of methanol
Figure 5.1

Figure 5.2
Comparison between biomass measurements as mg COD/g dry substratum and as
mg TS (TVS)/g ds,substratum in the wood chip biofilter during the biofiltration of methano1

Effects of Biumass Growth on Pressure Drop in Bioflters

Discussion

81

The detached biomass profiles (Figure 4.23), and TS and TVS profiles (Appendix H) in
both biofilters during the biofilation of a-pinene are very similar in each biofilter, indicating
that for this case these techniques suitably described biomass concentration in the beds. This
was probably due to a more homogeneous moisture content of the biomass in the biofilters

achieved by previous mixing, and due to lower microbial growth observed in this period, which
rnay have not affected the biomass composition. During this period, protein analyses were also
conducted in order to validate the volatile solids measurements. The protein profiles for the
Nova Inert biofilter (Appendix H) were qualitatively similar to those obtained using the other
techniques. During the biofiltration of a-pinene, the COD results for both biofilters and the
protein profiles for the wood chip biofilter showed a less uniform biomass distribution along the
beds (see Appendix H), probably due to changes in the composition of the biofilm and
interference fiom the decaying wood chips.
For purposes of funher analysis, just the profiles of detached biomass (g biolayedg dry
substratum) will be discussed since they were consistent with visual observations and pressure
drop measurements, and therefore, considered to most suitably characterise biomass
accumulation in the beds.
5.2 Pressure Drop and Cumulative Methanol Removal

When comparing the cumulative pressure drop profiles of both biofilters (Figures 4.12

and 4.13) with those of methanol concentration (Figures 4.3 and 4.4), the higher pressure drops
measured in the first top sections of the biofilters correspond to the greater cumulative methanol
removals also measured in these first two sections. The fact that most of the total pressure drop
in the biofilters was observed in the top sections, where most of the incoming methanol was
removed, suggests that methanol supported enough microbial growth and biomass accumulation
to occupy and/or clog void spaces within the bed, which caused an increase in airfiow
resistance. This behaviour was similar in both the Nova Inert and the wood chip biofilter.
Although both biofilters operated under similar removal efliciencies and methanol
loading rates, they generated different amounts of biomass and were subject to different
pressure drops. The pressure drop was on average dfold higher in the wood chip biofilter than

in the Nova hert bioflter (see Figure 4.1 l), consistent with greater biomass accumulation and
--

Eflects of Biomass Growfhon Pressure Drop in Bioflers

Discussion

82

compaction of the wood chip bed. Greater biomass accumulation in the wood chip biofilter
supports the idea that a naturai packing material such as wood chips, besides providing
nutrients, can aiso act as a second carbon source for microbial metabolism.
As an attempt to correlate the arnount of methanol removed and the pressure drop in the

biofilters, the pressure drop registered in each section at the end of each undisturbed-bed
operation period was plotted against the total amount of methanol removed in each section ai
that time. As can be seen in Figure 5.3 for the case of the Nova Inert biofilter, there exists a
non-linear relationship between pressure drop and methanol removed, with pressure drop
increasing more rapidly at high methanol removals. A similar, but less clear trend was obtained
for the wood chip biofilter (see Figure 5.4). In the wood chip biofilter, however, compaction
and the changes in the packing material infiuenced the correlation of pressure drop and biomass

accumulation.

Figure 5.3
Pressure drop dong each section as a function of the cumulative amount of
rnethanol removed in each section of the Nova Inert biofilter, measured at the end of each
undisturbed-bed operation period

Effects of Biomass Growth on Presmre Drop in Biofilers

Discussion

83

Figure 5.3 shows that there exists a critical value of methanol removal per bed section
(about 4000 gkection) after which pressure &op begins to increase steadily. This critical value
of methanol removal would correspond to a critical amount of biomass developed on the
medium that reduces porosity in such a way that airflow resistance increases significantly. This
is relevant to the operation of a biofilter since it suggests that pressure drop could be predicted
based on the amount of pollutant removed, and that there exists a critical removal amount before
pressure drop increases dramatically.

1000

2000

3000
4000
5000
6000
7000
Cumulative methand removed (glsection)

8000

9000

10000

Figure 5.4
Pressure &op along each section as a fnction of the cumulative amount of
methanol removed in each section of the wood chip biofilter, measured at the end of each
undisturbed-bed operation period

5.3 Pressure Drop and Biomass Accumulation during the Biofiltration of Methanol
5.3.1 Correlation between Pressure Drop and Biomass Accumulated in the beds
The biomass concentration profiles were also correlated with pressure drop profiles
along the beds. Higher pressure drops took place at the inlet sections of the beds, where most of

the methanol was degraded over t h e and higher biomass concentrations were measured. This

is illustrated by the results of Figures 4.5 and 4.6 for both biofilters. The detached biomass
Effects of Biomass Growth on Pressure Drop in Biofilrrs

Discussion

84

profiles clearly show that the greatest amount of biomass tended to accumulate at the top level

of the first section. The highest concentrations were measured at the end of period 2 (day 24Nova Inert; day 72-wood chips), when a thick biolayer formed at the top level of the first two
sections, especially in the wood chip biofilter. These high values of biomass in these sections
correlated with the highest pressure drops registered dong the beds, which were responsible for
almost 90% of the total pressure drop measured in this period in both biofilters (see Figure
4.1 1).
A plot of pressure drop per section versus amount of biomass accumulated in each

section of the biofilters at different times, as presented in Figure 5.5 for the Nova Inert biofilter,
gives also a non-linear, semiparabolic relation as for the case of cumulative methanol removal.
A similar relationship was obtained for the wood chip biofilter as s h o w in Figure 5.6. Cox et

al. (1998) obtained a similar parabolic increase of pressure drop with biomass concentration in

biotrickling filters removing toluene.

r Sedion 2
A Section 3 r

@section4

2 2000 a

g ,Of

E
&

1000 -

500
? - -O

Figure 5.5
Pressure drop dong each section as a function of the average amount of biomass
accumulated in each section of the Nova Inert biofilter, measured at the end of each undisturbed-bed
operation period

Eflects of Biomars Growth on Pressure Drop in Biofilers

Discussion

85

A steep increase in pressure drop takes place in the range of 3000 to 4000 g of biomass

for the Nova Inert biofilter, and in the range of 20 000 to 30 000 g of biomass for the wood chip
biofilter. These correspond to critical values of biomass accumulation, which mark the steady
increase in pressure &op. Figures 5.5 and 5.6 show that while in the Nova Inert biofilter
pressure drops under 500 Palm were registered before the sudden pressure drop increase at the
critical biomass values, in the wood chip biofilter, the pressure drops before the sudden increase
were below 1000 Pdm. This difference of twice as much pressure drop can be explained not
only by more biomass growth in the wood chip bed, but also by the degradation of the wood
chips and related bed compaction, and different bed pore structure. In addition, pressure drop in
the wood chip bed increased drastically when the amount of biomass accumulated on the bed
went over the mentioned critical value. This indicates that, although wood degradation and bed
compaction contributed to high pressure drops, it was the amount of biomass accumulated on
the wood chip bed that played the main role controlling the pressure drop increase. This was
clearly the case for the Nova Inert biofilter as shown by Figures 5.5 and 5.6.

*Section 1
iSedion 2
t

A Section 3
xSedion4,

Figure 5.6
Pressure drop dong each section as a function of the average amount of biomass
accumulated in each section of the wood chip biofilter, measured at the end of each undisturbed-bed
operation period

Effects of ~iomass~ r o w t hin Pressure Drop in Biojillers

Discussion

86

The results of detached biomass profiles show that biomass concentrations in the wood
chip biofilter were higher, nom 0.5 to 19 g biolayerl g dry substratum, than in the Nova Inert
biofilter, where concentrations fkom 0.4 to 3.2 g biolayerf g dry substratum were measured.
This was likely due to nutrient availability in excess of what was required for methanol
degradation, and to two substrate sources in the wood chip bed, that is, methanol and wood.
Higher biomass accumulation was believed to be the main cause of higher pressure drops
measured in the wood chip biofilter. Auria et al. (1993) also found that pressure losses in a
solid state fermentor packed with wheat bran and cane bagasse were controlled by mould
growth, even though these supports undenvent changes due to microbial growth. Compaction,
as a consequence of biomass growih and degradation of the wood chips also afTected pressure
drop in the wood chip biofilter; however, compaction did not seem to be the main factor
controlling pressure drop in the wood chip biofilter. As can be observed fiom Figure 5.7, for
the bottom sections of the wood chip biofilter (section 3 and 4), where low biomass growth took
place, pressure drop did not increase significantly even at the highest compaction values around
1.5 %. For sections 1 and 2, the high values of pressure drop do not correlate with percent

compaction (Figure 5.7), as did with biomass concentration. The complexity of segregating the
influence of different factors affecthg pressure drop in bioreactors packed with natural media,

has been recognised for solid state fermentors (Auria et al., 1993), and this is the case when
high biomass growth occurs simultaneously with compaction.
The link between biomass growth and methanol removal was frther investigated by
estimating a yield coefficient. For the Nova Inert biofilter a biomass yield coefficient of 0.01 g
dry biomass (TVS)/g rnethanol was obtained fiom the slope of the straight line of ploaing

biomass as g TVS versus methanol removed ( ~ ~ = 0 . 7 See


2 ; Appendix 1). This was not the case

for the wood chip biofilter where methanol and wood chips were used as substrate sources, and
there was low correlation between TVS and methanol removed (yield coefficient: 0.02 g TVS/g
methanol; ~ ~ 4 . 0 4 The
) . value of the yield coefficient for the Nova Inert biofilter is lower than
the values of yield coefficients for methanol degrading rnicroorganisms reported in the literature
(0.30-0.54) (Bailey and Ollis, 1986; Shareefdeen et al., 1993), probably due to a high
endogenous decay rate in the biofilter. It has to be kept in mind that these yield coefficients

Eflects of Biomas Growth on Pressure Drop in BioJiters

Discussion

87

reported in the literature were obtained for growing cultures in well mixed reactors and
endogenous decay is not considered.

Figure 5.7
Pressure drop as a function of percent compaction in each section of the wood chip
biofilter, measured at the end of each undisturbed-bed penod

53.2 Biomass Distribution Effects on Pressure Drop

Pressure &op axial profiles measured in both biofilters before and after mixing the beds
(Figure 4.12 and 4.13) illustrate the influence of biomass distribution on pressure drop in
biofilters. Different amounts of biomass and different pressure drops were measured in the beds
at the end of each undisturbed-bed operation period before remixing. Nevertheless, after
uniformiy distributhg the biomass by remixing the bed of each section, the total pressure drop
decreased significantly and almost to the same value around 40 Pa/m for the Nova Inert biofilter
and 100 Pdm for the wood chip biofilter. The decrease of pressure drop after remixing the bed

in each section indicates that pressure drops at the end of each period were much higher than
100 P a h because biomass was not eveniy distributed at that t h e before remixing. Biomass

tended to grow unevenly and accumulated more at the top level of the inlet sections, particularly

Effects of BiornPPs Growth on Pressure Drop in Bioflters

Discussion

88

at the end of penod 2, which caused signifiant pressure drop increases, probably by
channelling near the walls.
Lower pressure drops, compared with those at the end of period 2 and 3, were measured
at the end of the fourth undisturbed-bed operation period in both biofilters during the
biofiltration of methanol (Figures 4.12 and 4.13). This is explained by a more even biomass
distribution within and along the beds due to 3 previous remixings of each section, as s h o w by
the more uniform biomass profiles along the biofilters at the end of period 4 (Figures 4.5 and
4.6). These results also support the conclusion that an uneven biomass distribution (understood

as localised, high biomass concentrations) within the bed is the key factor afTecting increased
pressure drops in biofilten due to a non-optimal use of the void space within the bed. Thus, the
assessment of pressure drop effects due to biomass accumulation requires the consideration of
not only the arnount of biomass growing on the packing material, but also of its distribution on
it. Consequently, biomass distribution on the filter medium stands out as the factor upon which
the optimisation of the bed utilisation of a biofilter might rely, and it may suggest the need for
new designs of biofilters in order to achieve a uniform biomass distribution within the packing

bed. If lower pressure drops are achieved by even biomass distribution in the bed, compaction
might still have an impact on pressure &op since compaction would be caused by biomass
growth, and since the pressure drop wouid be low enough for compaction to have an important
effect.
5.4 Cornparison of Performance behveen Packing Media
5.4.1 Pollutant Removal

During the period of biofiltration of methanol and during the first 1O days of biofiltmtion

of a-pinene, both the Nova Inert and the wood chip biofilter performed similarly in tenns of
overall pollutant removal, under almost equal loading rates and EBRT's. Removal efficiencies
along the complete beds were above 95 % and corresponded to average removal capacities of
112 g methanol/m3bed in the Nova Inert biofilter and 127 g methanol/m3 bed in the wood chip
biofilter. The slightly greater value of the specific removal capacity in the wood chip biofilter
was a result of bed compaction, which reduced the bed volume.

Effects of Biomass Growth on Pressure Drop in Biofilers

Discussion

89

When comparing the removal capacities of the biofilters by sections as shown in Figure

5.8, it is observed that the Nova Inert biofilter tended to operate with slightly higher removal
capacities than the wood chip biofilter. However, the rernoval capacity in both biofilters seems
to reach a plateau around 100 and 140 g methanol/m3 bedh. The Nova Inert biofilter was
expected to have a higher removal capacity since the bed, composed of small porous pellets,
was believed to have a larger surface area per unit of bed volume than the wood chip bed. The
larger surface area was supposed to allow for greater biomass growth (Bryers, 1987) and mass
transfer, and given excess nutrients, a better removal per section and per total bed in the Nova
lnert biofilter. However, as seen fiom the methanol axial profiles (Figures 4.3 and 4.4) and
Figure 5.8, both biofilters tended to remove similar arnounts of methanol in each section, even
though the wood chip bed was suffering compaction. The exception was section 1 of the wood

chip biofilter, where a decrease in removal was observed after the first undisturbed-bed
operation period during the biofiltration of methanol.

NOMlnert biofilter
rn W o d diip bioiiter
removd rate = loading rate

--

Figure 5.8
Cornparison of the performance of the Nova Inert and wood chip biofilters in terms
of removal capacity in each section of the biofilters
The sirnilarity in methanol removal in both biofilters, in spite of a larger specific surface
area available in the Nova Inert bed, could be explained by the patchy pattern of biomass growth

Effects of Biumuss Growth on Pressure Drop in Biojlters

Discussion

aad accumulation in the beds.

90

In both biofilten the same patchy pattern of biomass

accumulation was observed, although higher biomass concentrations were obtained in the wood
chip biofilter, where the same amount of methanol was being degraded, but also wood
constinients could have been utilised as substrates. The formation a thick biomass layer at the
top levels seemed to be determined by the methanol loading rate in each section, which
decreased dong the bed as methanol was removed, and by the air residence time in the
biofilters, which was lower in the 4 bed sections placed in series, and higher in between each
section, where unpacked volume was available.
The decrease in methanol rernoval in the first section of the wood chip biofilter after the
first undisturbed-bed petiod of operation could have been caused by air channelling, probably
near the walls of the column, since a thick biofilm layer had fonned at the top level of this

section, which impeded airflow through it. Nonetheless, this was not the case for the last two
periods of operation since remixing the bed did not improve the performance in this section. A
constant, low pressure drop profile dong this section was observed just at the end of the fourth
undisturbed-bed period of operation; however, the amount of biomass in this section did not
decrease as s h o w by the detached biomass profiles (Figure 4.6). Since the sarne or greater
arnount of biomass was present in this section, but less methanol was being removed, the
substrate for the microorganisms could have corne from the wood chips.

The performance of the wood chip biofilter began to detenorate around day 160, after 10
days fiom switching over to a-pinene. The removd efficiency dropped from 100 to 40%
removal under an inlet loading rate of 10 g cepinene/m3bed h in a period of 45 days (see Figure
4.21). During the biofiltration of a-pinene, the wood chips becarne dark-brown and they

seemed to be covered with white fhgi. In contrast, the Nova Inert biofilter remained operating
with 100% removal efficiencies for a longer period, until day 190, even d e r remixing the bed
and experiencing overheating at 55 OC during 5 h around day 155, which required the system 4
days to achieve 100% removal again. The robustness of the Nova Inert biofilter as compared
with the wood chip biofilter may lie on the inert characteristics of the packing material, which
did not compte with the degradation of methanol and a-pinene.
The decrease in removal efficiencies observed in both biofilters around day 190, could
have been related to a depletion in nutrients since just macronutrients were added to the Nova
Eflects of Biomass Growth on Pressure Drop in Biofiters

Discussion

91

Inert biofilter during bed remixing before switching over to a-pinene. This is in agreement with
observations made in previous biofiltration experiments using a-pinene and the same fertiliser
pellets, where a lower performance of the biofilters was related to low nutrient availability in the
bed around 150 to 200 days of operation (Mohseni, 1998). It is possible that nutrients were first
depleted in the wood chip biofilter since more carbon as methanol and wood would have been
consumed.
5.4.2 Pressure Drop

The moisture content of the bed is another important factor that would have affected
pressure drop in the biofilters. Bed moisture contents varying fiom 25 to 60% in weight were
measured in the Nova Inert bed, and fiom 50 to 80% in the wood chip bed.

Biomass

quantification as detached biomass (g biolayer/ g dry substratum) took into account the bed
moisture content since the biomass quantified corresponded to the difference between
substratum with wet biomass and dry substratum without biomass. However, detached biomass
concentrations in the wood chip biofilter were higher not only because of the water content, but
also because of the greater amount of biofilm, as shown by the higher TS and TVS content of
the biomass sarnples in the wood chip biofilter as compared with that of the Nova inert biofilter.
Nevertheless, moisture content seemed to strongly influence pressure drop in the biofilters. In
the wood chip biofilter, where higher pressure drops were measured, aiso a high bed moisture
content was registered, as compared with an optimal moisture content of 30 to 60% by weight
(Williams and Miller, 1992b). The higher pressure drops in the wood chip bed observed since
the start-up period c m be related to the high moisture content of the bed, however moisture
content per se was not responsible for the steady increase in pressure losses. Higher water
content in the wood chip bed, besides occupying void space within the bed and reducing bed
porosity, may have contributed to greater biomass growth and accumulation, and to an
accelerated wood chip degradation. The increase in moishm content in each section of the
biofilters over tirne (Figures 4.17 and 4.18) was correlated to an increase of biomass in each
section over time, which suggests that the moisture content in the bed depended on the amount
of biomass present.

Eflects of Biomoss Growth on Pressure Drop in Bioflters

Discussion

92

5.4.3 Biomass Accumulation

During the first two undisturbed-bed periods of operation with methanol, biomass
accurnulated at the top levels of the first two sections of both biofilters, but also near the walls
of the column in each section, probably due to biomass sloughing fiom the thickest portions.
High loading rates in the first sections of the biofilters could have been the cause of the

formation of these biomass layers. As discussed previously, greater biomass accumulation was
observed in the wood chip bed. The development of a continuous biomass layer has been
reported in wastewater biofilm systems, and it is believed to originate fiom ce11 microclusters
with extracellular polymeric substances (EPS) in the voids (Bishop, 1997). This could have

been the case at the top levels of the first sections of the biofilters, where high methanol
loadings allowed the microclusters and the continuous layer to develop.
Biomass sloughing, even though not well studied, has been reported to occur in biofilm
systems, and may play an important role in biofilm formation and persistence (Bryers, 1987). In
the centre of the sections of both biofilters, where dry regions were detected, hingal growth was
observed. Biomass growing on the moist parts of the bed was responsible for the increased
pressure drop in the biofilters, and as revealed from its analysis ( F w g i a , 1999), had a high
carbohydrate content, which suggests that the microorganisms were degrading methanol and
storing it as carbohydrates in the form of EPS. The EPS could also have supported a population
of non-methanol degrading organisms as suggested by other authors in biofiltration experiments
(Smith et al., 1996).
It is hypothesised that a mon even biomass distribution or less localised high biomass
concentrations could be achieved fiom the start-up period, without remixing the bed, if the
biofilter is fed with a uniform VOC loading rate along the biofilter. The initial maner of how
biomass accumulates on the packing materiai may have an impact on its later development.

This would allow for a better airflow distribution due to the development of a microbial
community subject to the same substrate loading rates, without regard to their location along the

column.

This would be analogous to a submerged rotating drum reactor, where under

completely mked conditions, unifonn biofilm growth is achieved (Zhang et al., 1997).

Effects of Biomms Growth on Pressure Drop in BiofIters

Discussion

93

5.5 Airflow characteristics


5.5.1 Correlation between Pressure Drop and Airfiow rate
The results from plotting pressure drop as a fnction of airflow rate (Figures 4.14 and 4,
15) showed that biomass accumulation over time increased the pressure &op in the biofilters;

however, the increase was larger at the highest airflow rates (up to 130 Llmin). At the highest
flow rates, which were close to the actual airflow rates of operation (around 120 L/min), the
increase in pressure &op in the plots of AP versus flow rate, was described by an increase in the
degree of curvature of the lines. This indicates that the increase in pressure drop was due to an
increase in both kinetic and viscous energy losses, where kinetic energy losses becarne the
controlling losses rather than viscous energy losses, which are the controlling factor at low

airflow rates. An examination of the modified Ergun equation (Eq. 2) supports this observation
since the term representing kinetic energy is a function of the flow rate squared, which is
responsible for the parabolic shape of the curves.
Figure 4.14 and 4.15 also suggest the direct effect of biomass accumulation on the
increase of pressure &op. More biomass accumulation reduces the porosity of the bed, which
causes both viscous and kinetic energy losses to increase; however, the kinetic losses increase
more rapidly than the viscous losses as s h o w by the increase of the degree of curvature of the
lines over tirne. This is explained by the more rapid increase of the factor (~-E)/E"~ in the term
of kinetic energy losses with decreasing porosity

(E)

than the factor ((1-E) 2)/E

3.6

in the term of

viscous energy losses in the Modified Ergun equation.


Biomass growth affects the aimow characteristics through the bed by increasing more
significantiy the kinetic energy losses than the viscous energy losses. M i l e viscous energy
losses are predorninant in laminar flow, kinetic energy losses are characteristics of turbulent
flow (Ergun, 1952; Macdonald, et al., 1979). This knowledge coupled with the Reynolds
numbers for flow through packed beds ((NR~)~)
for these experirnents (Appendix J), which were
around 2 to 14 for the Nova Inert biofilter, and in the range of 2 to 11 for the wood chip
biofilter, indicate some characteristics of a turbulent regime. Even though for (NRc)b 2 larninar
flow prevails (Holland, 1973), there is controversy in estimating a transition Reynolds number
between laminar and turbulent flow in porous media (Kuwahara et al., 1998). Hence, whether
Effects of Biomass Growth on Pressure Drop in Biofiters

Discussion

94

laminar or turbulent flow regirne prevailed is difficult to assess; however, it cm be considered


that, given the increase in kinetic energy losses and the change of curvature of the lines of AP
versus flow rate at high flow rates, a transitional flow regime could have predominated. Slightly
greater Reynolds nmbers were measured during the first undisturbed-bed penod of operation in
both biofilters; (NRe)b
as high as 19 were calculated for the Nova Inert bed, and as high as 13 for
the wood chip bed. This corresponded to the formation of the thick biofilm at the top the first

two sections, and suggests characteristics of a hubulent flow regime in the beds, as suggested by
the Reynolds numbers and kinetic energy losses.

Similar deviations from linearity of AP versus airflow rate curves in biofilters have been
reported in other biofiltration studies (Leson and Winer, 1991;Sabo, et al., 1993).
5.6 Modelliig Pressure Drop vs. Biomass Accumulation

The Ergun equation c m be used to obtain a rough estimate of pressure drop in biofilters,

as suggested by Ottengraf et al. (1986), and, in the sarne way, the Modified-Ergun (M-E)
equation for rough particles (Macdonald et al., 1979) could be used to estimate pressure drop in
biofilters. However, predicting air pressure losses in a biofilter still remains a challenge since
the variations in bed moisture content, biomass growth, and compaction affect the porosity of
the bed in such a marner that great differences between the predicted and expenmental pressure

&op &ta may arise.


It is believed that the success of the M-E equation for predicting pressure drop in a
biofilter relies on the success of predicting bed porosity changes through the biofilter's

operation period. For an inert packing material such as the Nova Inert pellets, since there is no
compaction, biomass growth plays the main role decreasing porosity and increasing pressure
drop across the bed. For the case of n a d packing materials, predicting biomass growth
effects on bed porosity represents the first step that needs to be taken in order to obtain a mode1
that predicts bed porosity changes due to biomass growth coupled with aging phenornena such

as compaction.

--

Effects of Biomass Growth on Pressure Drop in Biofiters

Discussion

95

5.6.1 Model Development

As a first approach to predicting airfiow pressure losses due to biomass growth in


biofilters, an expression used to calculate porosity changes in packed beds and the M-E equation
are used to estimate pressure drops in the Nova Inert biofilter. The phenomenological mode1
proposed here estimates pressure drop based on airflow rates and on the experimental
concentrations of biomass (g biolayed g dry substratum) taken at different bed levels and at the
end of each undisturbed-bed operation period in the Nova lnert biofilter during the biofiltration
of methanol. The following assumptions are considered:
1. The packing material is composed of inert spheres with the same average diameter.
2. Biomass grows uniformly on the surface of the spheres; and the spheres remain in

contact during biomass accumulation, which does not consider biofilm growth on the
whole surface of the spheres.
3. Biomass has a constant density within the biofilm, which can be assumed at a macro

scale, even though at a micro scale there are strong evidences of high biofilm
heterogeneity (Bishop, 1997).
The initial data required are:

1. Bulk density of dry substratum (g dry substratum/ m3 dry substratum), determined


experimentally.
2. Initial bed porosity (E.), determined experimentally or estimated.

3. Shape or sphericity factor of the packing material, which for the Nova Inert pellets is

assumed to be 0.85, after consulting values of 0.83 for rounded sand, and of 0.75 for an
average of various sand types (Perry and reen, 1988).
4. Biofilm density, which was determined experimentally.
5.6.1.1 Estimation of Biofilm Thickness

In order to take into account the decrease in bed porosity due to biomass accumulation,
biofilm thickness values are first estimated fiom biomass concentrations (g biolayedg dry
substratum) in the bed. The biomass concentrations are converted fiorn g biolayer (biomass)/g
dry substratum to g biomasslm2substratum surface using the dry substratum buk density and an
estimated initiai superficiai area of the packed substratum (Appendix M). The initial surface
Effects of Biomass Growth on Pressure Drop in Biofilers

Discussion

96

area per'volurne of clean substratum (m2/m3 ), a , is estimated fiom the following expression
(Perry and Green, 1988):

where 4 is the sphericity factor for the packing solids, and R is the radius of the sphere
equivalent to the packing medium, in this case, considered to be the average radius of the
pellets. Then, the biofilm thickness (LI) is calculated as:

where X is the superficial biomass concentration in g biornass/ m2 surface area, and p~ is the
biofilm density in g biornass/m%iomass.

5.6.1.2 Estimation of Biomiss-affected Porosity


Once the biofilm thickness has been calculated, the porosity of the bed with biomass
development can be estimated. For this, the equation for biomass-affected porosity proposed by
Alonso et al. (1997) is used. This equation was deduced for spheres in contact, upon which
biofilm grows uniformly, leaving the contact points uncovered. The equation is as follows
(refer to Eq. 4 in section 4.0):

This equation requires the coordination number, n, which accounts for the number of
contact points among the spheres.

This number can be assumed or estimated fiom an

expression correlating porosity and the coordination number for a close randorn type of packing
(Dullien, 1992):

With an initiai porosity of 0.52, deterrnined experimentally, the two solutions of Eq. 3 1

are: nr=22 and n2=6. Since the value of n ranges fiom 6, for a cubic arrangement, to 12, for a
rhombohedral arrangement, the value of n=6 is selected.

Effects of Biomass Growth on Pressure Drop in BiofLters

Discussion

97

Other biomass-affected porosity expressions were also tested.

These were the

expression deduced by Cunningham et al. (1991) for spheres packed in a cubic arrangement
(refer to Eq. 5 in section 2), and the expression obtained by (Taylor et oz., 1990) for porosity
with a biofilm thickness for a heterogeneous size distribution of spheres (refer to Eq. 12 in

section 2). While Cunningham's expression (Eq. 5) gave negative values for the experimental
biomass concentrations used, indicating complete clogging of the void space, Taylor's equation
(Eq. 12) did not predict any change in porosity based on the diflerent experimental values of

biomass concentration. In contrast, the porosities calculated using Alonso's expression (Eq. 4)
decreased with increasing biomass concentration, and were close to values of porosity
determined experimentally.
5.6.1.3 Estimation of Pressure Drop

Pressure drop across the Nova Inert bed was estimated using the Modified-Ergun
equation (refer to Eq. 2 for more details) proposed by Macdonald et al. (1 979). The use of the

M-E equation is justified by its consideration of porosity as a variable that characterises the
packing medium, and its effective application for describing pressure drop in compost-wood
chip mixtures (MacFarlane, 1998).
The coefficients used in the MO energy losses terms of the M-E equation were A= 180
for the viscous losses, and B=4 for the kinetic losses. The coefficient B=4 was used since it is
recommended for particles with high surface roughness (Macdonald et ai., 1979), and the Nova
Inert pellets had a rough surface which was considered to become rougher due to uneven
biomass accumulation as observed in these experiments. Pressure drop was then evaluated at
each biomass-aected porosity, estimated previously.
Pressure drop was also estimated fiom values of biomass-affected, specific permeability,
which were estimated using Equation 20 as proposed by Clement et al. (1996), and the equation
used by Auria et al. (1993) to predict specific permeability in a solid state fermenta-:

where k is specific permeability, E is porosity, Dp is the average particle diameter, and t is the
tortuosity factor (1.58 for sphens). Equation 32 is equivalent to the capillaric permeability
Effects of Biomass Growth on Pressure Drop in Biofiters

Discussion

98

model known as Carman-Kozeny equation (Dullien, 1992). The values of biomass-affected,


specific permeability were substituted in Darcy's law (Eq. 7) and pressure drop was calculated.
The pressure drop values, when compared to the experimental ones and to the results obtained
using the M-E equation, were very low, and hence were not considered for firther analysis.
5.6.2 Comparison between Experimental and Estimated Pressure Drop

In order to compare the estimated values of pressure drop with the experimental data
fiom the Nova Inert biofilter, both series of pressure drop data were plotted as a function of the
experimental biomass concentration as g biolayed g dry substratum, as presented in Figure 5.9.
Figure 5.9 shows the experimental AP measured in every subsection of each section (recall that

AP was measured in 2 subsections of each section of the biofilter), and the AP estimated by the
model. Both series of AP data are plotted versus the corresponding biomass concentration at the
top of each section for the AP of the first subsection, and the average of biornass concentrations
at the middle and bottom of each section for the AP of the second subsection.
It can be observed in Figure 5.9 that the estimated pressure drops lie under the

experimental values, suggesting an underestimation by the proposed model. This is due partly
to the biomass measurement technique that averages biomass concentrations over a specific
volume. However, one assumption of the model is that biomass accumulates uniformly on the
packing material, and as explained in previous sections, this was not observed in these
experiments, hence, these results are not suprising. This difference in results leads to identiQ
two important issues. First, that the patchy and uneven biomass accumulation (localised high
biomass concentration versus low biomass concentration in some parts) in the biofilter bed
caused a more pronounced increase in pressure &op rather than a smooth increase as proposed
by the model. And second, that if biomass wodd have accumulated uniformly on the Nova

Inert pellets, the pressure drop would not have reached the highest values of close to 3000 Pajrn,
registered in the top section of the bed. This last observation implies that if the uneven biomass
distribution could be taken into account in the prediction of the model, e.g. by measuring
differential spatial distribution of biomass concentration in the bed, then the pressure drop
values predicted wouid be higher and closer to the experimental values. This shows the

Effects of Biomass Growth on Pressure Drop in Bio$ilters

Discussion

99

relevance of achieving a uniform biomass distribution in the bed, which would allow for lower
pressure drops during the operation of a biofilter.

Experimental and estimated by the proposed model pressure drop values for the
Figure 5.9
Nova Inert biofilter during the biofiltration of methanol
As part of a frther analysis of the model developed in this section, bed porosity values
were calculated fkom experimentd values of pressure drop (2 AP values per section) using the
Modified-Ergun equation. These caiculated porosities and the porosities estimated using the
model for biomass growth on spheres in contact (Eq. 4) from experimentd values of biomass
concentration (same values used as in Figure 5.9) were plotted as presented in Figure 5.10.

Figure 5.10 also shows the values of wet porosity (detemillied by filling the bed void space with
water) measured fiom sarnples taken fiom each mixed section of the Nova Inert biofilter, as a
function of the average biomass concentration per section, and as a function of pressure drop.

Effects of Biomars Growth on Pressure Drop in Biofi/ters

Discussion

1 00

g Udayerlg dry subetratum


0.6
-

6 ~odified-~rgun
equation

Figure 5.10 Comparison between wet porosity measured fiom well mixed bed samples and
the bed porosity in the Nova lnert biofilter as predicted by the M-E equation fiom AP
expenmental data, and by the biomass growth model on spheres in contact with each other
There is a gap between the estimated porosities fiom the biomass growth model and the
calculated porosities from the M-E equation. This is not surpnsing since the growth-mode1
series was estimated assurning even biomass distribution on the spheres and dong the bed
sections for which the experimental biomass concentrations were used (top and average of
rniddle and bottom). On the other hand, the M-Eequation series was calculated fiom actual
experimental values of pressure drop registered for uneven biomass accumulation; however, the

M-E equation inherently predicts much lower bed porosities based on the AP values. This
shows that biomass growth, when occurring in an uneven way, controls more significantly
pressure drop in a biofilter. And in order to predict values of pressure drop in agreement with
the experirnental rneasurements, it is necessary to consider not only biomass concentration, but
also its distribution on the packing material with its consequent effects on bed porosity. Thus,
the difference in the predicted porosities may be due to uneven biomass growth, which caused

higher pressure drops, rather than due to lower pressure drops as for the case of a more even
biomass growth.

Effects of Biorners Growth on Pressure Drop in Biojiters

Discussion

101

Figure S. 10 also shows that the porosity values predicted by the biomass-growth model
are in good agreement with the experimental values of wet porosity measured fiom sarnples
taken fkom each section of the Nova Inert biofilter, afler mixing. In contrast, the wet porosities

as a function of pressure drop do not present the sarne agreement with the porosities predicted
fiorn the M-E equation due to the effect of uneven biomass distribution on high values of
pressure drop. In conclusion, the mode1 for estimating bed porosity based on biomass growth
on spheres in contact, adequately predicts the porosity changes due to even biomass
accumulation, and is a good estimate for predicting biomass-affected porosity in beds packed

with sphencal-like pellets.


5.6.3 Cornparison among Models

The three series of pressure &op data predicted fiom the experimental values of biomass
concentration using the models (refer to Appendix M for detailed calculations) are lower than
the experimental ones (see Figure 5.1 1). Figure 5.1 1 illustrates that pressure drops estimated
using specific permeability rnodels and Darcy's law are significantly lower than those calculated
fiom the Modified-Ergun equation. The underestimation by the specific permeabilityDarcy's
models might mise fiom the assumptions of these models. The macroscopic model proposed by
Clement et al. (1996) considers functions that describe the distribution of pore radii that is
derived from Van Genuchten's and Brooks and Corey's functions, which are soil-water

(drainage) retention relations for descnbing the relationship between relative water saturation
and pressure head in porous media. One source of error could be the consideration that the pore
size distribution function rernains constant even with biomass growth as assumed by Clement's
permeability model.

In addition, pressure &op underestimation might also arise fiom

considering that only viscous energy losses affect pressure drop in the biofilter, as stated by
Darcy's law, instead of considering both viscous and kinetic energy losses, the last of which
become more significant with biomass growth as shown by the results of this study.

Discussion

102

Figure 5.1 1 Cornparison between values of pressure drop estimated by the explained models
and the experimental values fiom the Nova Inert biofilter

The estimated values of pressure drop as presented in Figure 5.9 are obtained from
experimental values of biomass concentration (g biolayed g dry substratum). The biomass
concentrations used correspond to the concentration at the top of each section, and the average
value of the concentration at the middle and bottom of each section at the end of each
undisturbed-bed period before remixing the bed (see Appendix M).

Since these biomass

concentrations describe a profile dong the bed with the highest biomass concentrations at the
inlet sections, the mode1 predicts the highest pressure drops at these biomass concentrations as
well; however, not as high as the experimental values, due to averaging biomass concentration
over a specific bed volume and due to uneven biomass distribution. Part of the underestimation
could also originate fiom inadequately measuring biomass concentration spatial distribution in
the bed.

On the other hand, when pressure drops are estimated fiom the average biomass
concentration in a section (average of biomass concentration at the top, rniddle, and bottom of
each section), the estimated pressure drop values are much lower than the predicted data shown

in Figure 5.9. This happens because the predicted pressure drops correspond, for this last case,
Eficts of Biomass Growth on Pressure Drop in Biofilrers

Discussion

103

to lower biomass concentrations obtained fiom averaging the top, middle, and bottom biomass
concentrations in each section. This average biomass concentration could be considered as
values of biomass concentration for a unifonn biomass growth on the Nova Inert pellets in each
section. In order to prove this hypothesis, the pressure drop measurements after remixing the
bed were plotted against biomass concentration in the same graph as the estimated pressure
drops fiom the averaged (top, middle, bottom) biomass concentrations (see Figure 5.12). Figure

5.12 shows that there is a good agreement between these two series of data, indicating that the
model successfully predicts pressure &op for the case of even biomass growth on the packing
material, and that deviations from the experimental data before remixing (as in Figure 5.9) are
due to uneven biomass distribution at high biomass concentrations. This also verifes that the
success of the model lies on the prediction of biomass-affected porosity, coupled with the
Modified-Ergun equation for rough particles, which takes into account the viscous and kinetic
energy loss increase due to biomass accumulation.

0.0

Experirnentalafter remixing
Experimental befm remixing j

0.5

1.O

1.5

2.0

2.5

g biolayert g dry substratum

Figure 5.12 Cornparison between estimated pressure drop from averaged biomass
concentrations and experimental pressure &op data before and &er remixing the Nova Inert bed

Conclusions and Recommendations

104

6 CONCLUSIONS AND RECOMMENDATIONS


1. Uneven biomass distribution, described by localised high and low biomass concentrations

within the biofilter bed, is the key factor affecting increased pressure drops in biofilters due
to local clogging andor extreme local void space reduction. The optimal utilisation of a
biofilter bed strongly depends on the amount of biomass growing on the packing material,
but also on its distribution on it.

2. The highest pressure drop in the beds was caused by thick layers of biomass with high
extrapolymeric substance concentration and high moishve content, both of which increased
the biomass specific volume and significantly decreased the bed porosity at the top levels of
the first sections of the biofilters.
3. A non-linear, semiparabolic correlation exists between pressure drop along the biofilter and

amount of biomass accurnulated in the bed, and cumulative methanol removal. There is a
critical value of methanol removal, and of biomass accumulation, after which airfiow
resistance increases significantly. This is relevant to the operation of a biofilter since it
suggests that pressure &op could be predicted based on the arnount of pollutant removed.
4. Wood chips caused higher pressure drops, on average 6 times higher, than the Nova Inert

pellets. Compaction, as a consequence of biomass growth and wood chip degradation,


together with moisture content increased pressure drop in the wood chip bed.

5. Biomass accumulation on the wood chip bed, compared with compaction, piayed the main
role controlling the pressure &op increase. Wood chips could have enhanced nutnent
availability and provided a second carbon source for microbial metabolism, which caused
greater biomass accumulation. It is recommended that wood chip degradation and its relation
to nutrienthubstrate consumption and biomass growth in the bed should be M e r
investigated in order to understand these phenomena and operate biofilten using wood chips

as packing materials more efficiently.


6. A mode1 was proposed that successfully predicts pressure &op for the case of even biornass
growth on sphencal particles. Deviations fiom experimental data before remixing are due to
uneven biomass distribution at high biomass concentrations, which is not thoroughly
characterised by averaging biomass concentration over a specific bed volume. The success

Eflects of Biomass Growth on Pressure Drop in Biofilrers

Conclusions and Recommendations

105

of the model lies in predicting biomass-affected porosity, coupled with the Modified-Ergun
equation for rough particles, which takes into account both viscous and kinetic energy losses
increase due to biomass accumulation.
7. The model of biomass growth on spheres in contact with each other predicts higher bed

porosities compared to porosities predicted by the Modified-Ergun equation based on


experimental pressure drops due to averaging biomass concentration over a specific bed
volume. But also due to the inability of this model to predict lower porosities due to uneven
biomass accumulation. This mode1 is a good estimate for predicting biomass-affected
porosity in beds packed with sphencal pellets where biomass grows evenly distributed.
8. The underestimation of pressure drop by the specific permeability models and the Darcy's

law arise likely fiom averaging biomass concentrations over a bed volume, and fiom
considering that the pore size distribution function involved in Clement's permeability model
remains constant with biomass growth. Further underestimation is due to considering that
only viscous energy losses affect pressure &op in the biofilter, as stated by Darcy's law,
instead of considering both viscous and kinetic energy losses, as in the Modified Ergun
equation.
9. Further studies are recommended to evaluate the impact of unifonn VOC loading rates on

even biomass distribution along a biofilter, which could imply a distribution of the incorning
air dong the biofilter bed in a new biofilter design. Given a uniform air distribution, the
initial uniforni pattern of biomass accumulation in the bed may ensure a more even biomass
distribution throughout the operation period.
10. Composite sampling at the top, rniddle, and bottom of each section and biomass

measurement by the detached biomass technique (g biolayedg dry substratum) indirect1y


quantified the volume of biomass accumulated on the media, and described its uneven
distribution within and along the beds. The COD tests and the TSiTVS tests suitably

described biomass concentrations in the bed responsible for pressure drop, except when
biomass had a high water content, as at the top of the fmt sections of the biofilters.
1 1. Biomass accumulation in the biofilter bed reduces the porosity of the bed, which causes

both viscous and kinetic energy losses to increase. While at lower superficial gas velocities

Effects of Biomass Growth on Pressure Drop in Biofilers

Conclusions and Recommendations

106

pressure &op is controlled by viscous energy losses, at higher superficial gas velocities,
kinetic energy losses become the controlling factor.
12. It is recomrnended that tracer tests in biofilters be conducted in such a way that any

interference from the equipment quantifying the tracer concentration be able to be

accounted for. This could involve tracer step changes, instead of tracer pulses, and the
measurement of the tracer concentration at the inlet and outlet of the biofilter. In addition,
smoke tests and radiai measurements of the pollutant concentration at different bed levels
could help to gain insight into the chatacteristics of the flow through the biofilters.

E m s of Biomass Growth on Pressure Drop in Biofiters

7 NOMENCLATURE
surface area per volume of clean substratum
Constant in viscous energy loss term in M-E equation
cross sectional area
Constant in kinetic energy loss term in M-Eequation
constant equal to 115
sphere diameter
equivalent diameter of packing particles
average particle diarneter of the clay balls
effective diameter of particles
pore size distribution fiuiction
acceleration due to gravitgravitational constant
air mass flow rate
piezometric head
specific pemeability
biofilm-affected specific pemeability
apparent first order rate constant
hydraulic conductivity
biofilrn thickness
specific surface
biomass-afYected specific surface
biomass-unafXected porosity (Taylor's model)
nurnber of packing spheres in contact with a single sphere
biomass-af5ected porosity (Taylor's model)
biomass volume fraction (volume of biomass/totai volume)
Reynolds number for flow through the biofilter beds
water flow rate
pore radii of a pair of serially connected capillary elements
minimum and maximum p r e radii
radius of the sphere equivalent to the packing medium
gas Reynolds number
liquid Reynolds number
outlet and inlet substrate concentration in the gas
Effects of Biomuss Growth on Pressure h o p in Biofilers

Nomenclature
t

u,
VSG
VSL
x
X

x'

108

tortuosity factor
residence time in the bed
superficiai air velocity
superficial velocity of the gas
superficial liquid velocity
distance dong the column
superficial biomass concentration g biomassl m2 surface area
dry biomass weight per unit mass of aquifer solids

Greek Letters

packing arrangement factor ( m = number of contact points)


dimensionless constants
pressure drop across the bed of height L
bed fiactional void volume or porosity
initial porosity of the bed without biofilm
biornass-dected porosity
fluid density
solid-phase biomass or biofilm density
gas density
bulk density of aquifer solids
moisture content fiinction
dimensionless constants
fluid absolute viscosity
sphericity factor for the p a c h g solids

Pa n i '

Abbreviations

ATP
CFU

COD
EBRT
(U.S.) EPA

EPS
GAC

adenosine triphosphate
colony forming units
chernical oxygen demand
empty bed retention tirne
United States Enviromentai Protection Agency
extracellular polymeric substances
grandar activated carbon
Effects of Biomass Growth on Pressure Drop in Biof fters

Nomenclature

GC

1O9

gas chromatograph

hazardous air pollutants


maximum achievable control technology
Modified Ergun equation
methyl ethyl ketone
MEK
methyl isobuthyl ketone
MIBK
RH (as in 99% RH) relative hurnidity
soiid state fennentor
SSF
total hydrocarbon analyser
T m
therrno-mechanical pulping
TMP
total solids
TS
total volatile solids
TVS
volatile organic compound
VOC

HAP
MACT
M-Eequation

Effects of Biomuss Growth on Pressure Drop in BioJiters

8 REFERENCES
Allen, E.R. and Y. Yang (1991). "Biofiltration control of hydrogen sulfide emissions". 84th Air
& Waste Management Association Annual Meeting & Exhibition, Vancouver, British
Columbia, June 16-21.
Allen, P. J. and T. S. Van Ti1 (1996). "Installation of a full scale biofilter for odor reduction at a
hardboard mill". Conference on Biofiltration (an Air Pollution Control Technology),
University of Southem California, Loa Angeles, October 24-25,3 1-43.
Alonso, C., M. T. Suidan, G. A. Sonal, F. L. Smith, P. Biswas, P. J. Smith, and R. C. Brenner
(1997). "Gas treatment in trickle-bed biofilters: Biomass, How much is enough?".
Biotechnolqp and Bioengineering, 54(6): 583-594 pp.
Andreoni, V., G. Origgi, M. Colombo, E. Calcaterra, and A. Colombi (1997). "Characterization
of a biofilter treating toluene contaminated air". Biodegradation 7(5): 397-404 pp.
Apel, W.A., B. D. Lee, M. R. Walton, L. L. Cook, and K. B. Dinerstein (1995). "Removal of apinene fiom off-gases by biofiltration". 88th Annual Meeting & Exhibition, San Antonio,

Texas, June 18-23.


APHA (1992). "Standard Methods for the Examination of Water and Wastewater". 18 ed,
Washington, D. C., Amencan Public Health Association.

Auria, R., M. Morales, E. Villegas, and S. Revah (1993). "Influence of mold growth on the
pressure drop in aerated solid state fermentors". Biotechnology and Bioengineering,
41(11): 1007-1013 pp.
Auria, Rny 1. Ortiz, E. Villegas, and S. Revah (1995). "Influence of growth and high mould
concentrations on the pressure drop in solid state fermentations". Process Biochernistry,
30(8): 751-756 pp.
Bagley, D. M. (1997). "CN 1309s Laboratory #2 hand-out". University of Toronto.
Bailey, J. E. and D. F. Ollis (1986). "Biochemical engineering fhdarnentals". 2nd ed, New
York, McGraw-Hill, Inc.
Bardtke, D., K. Fischer, and F. Sabo (1987). "Air purification with biofilters, field of application
and design cnteria". 80th Annual Meeting of APCA, New York, New York, June 21-26,
1007-1013.
Bishop, P. L.(1997). "Biofilm structure and kineticsl'. Wat. Sci. Tech., 36(1): 287-294 pp.
Effects of Biomass Growth on Pressure Drop in Biojlters

Bryers, J. D. (1987). "Biologically active surfaces: Processes governing the formation and
persistence of biofilms". Biotechnology Progress, 3(2): 57-68 pp.
Characklis, W. G. (1990). "Laboratory biofilm reactors" in Biofilms. W. G. Characklis and K.
C. Marshall, New York, U.S.A., John Wiley & Sons, Inc. 73-87pp.
Cherry, R.S. and D. N. Thompson (1997). "Shift fiom growth to nutrient-limited maintenance
kinetics during biofilter acclimation". Biotechnology and Bioengineering, 56(3): 3 30-33 9

PP*
Clement, T. P., B. S. Hooker, and R. S. Skeen (1996)."Macroscopic models for predicting
changes in saturated porous media properties caused by microbial growth". Ground
Water, 34(5): 934-942pp.
Coleman, R. N. and E. C. Dombroski (1996). "Development of biofilters to scrub emission
compounds fiom wood processing plants", Alberta Economic Development and Tourism.
Cooksey, K. E. (1 992). "Extracellular polyrners in biofilms" in --Science
and
.L. F. Melo, T. R. Bott, M. Fletcher and B. Capdeville, The Netherlands,
Kluwer Academic Publishers 1 37- 147 pp.
Corsi, R. L. and L. Seed (1995)."Biofiltration of BTEX: Media, substrate, and loadings effects".
Environmental Progress, 14(3): 1 5 1-1 58 pp.
Cox, H. H. J., T. T. Nguyen, and M. A. Deshusses (1998)."Elimination of toluene vapors in
biotrickling filters: perfomiance and carbon balances". Air & Waste Management
Association's 91st Annual Meeting & Exhibition, San Diego, California, June 14- 1 8,
paper 98- WAA.04P.

Cunningham, A. B., E. J. Bouwer, and W. G. Characklis (1990). "Biofilms in porous media" in


.
Biofilms. W. G. Characklis and K. C. Marshall, New York, U.S.A, John Wiley & Sons,
Inc. 697-732pp.
Cunningham, A. B., W. 0.Characklis, F. Abedeen, and D. Crawford (1991)."Inlfbence of
biofiim accumulation on porous media hydrodynamics". Environ. Sci. TechnoL, 25(7):
1305-1311 pp.
De Castro, A., D. G.Allen, and R. R. Fulthorpe (1997). "Characteriuition of the microbial
population during biofiltration and the infiuence of the inoculum source". Air and Waste
Management Association's 90th Annual Meeting and Exhibition, Pittsburgh, PA.
Dence, C. W. and D. W. Reeve (1996). "Pulp bleaching. Pnnciples and practce", Atlanta,
TAPPI Press.

Eflects of Biomass Growth on Pressure Drop in Biofllters

Deront, M., F. M. Samb, N. Adler, and p. Pringer (1998). "Biomass growth monitoring using
pressure drop in a cocurrent biofilter". Biotechnology adn Bioengineering, 60(1): 97- 104

PP*
Deshusses, M. A. and H. H. J. Cox (1998). "The use of CAT Scanning to characterize
bioreactors for waste air treatment". Air & Waste Management Association's 91 st Annual
Meeting & Exhibition, San Diego, California., June 14-1 8, paper 98-TA20B.04.

Deshusses, M. A., G. Hamer, and 1. J. Dunn (1995). "Behavior of biofilters for waste air
biotreatment. 1. Dynamic mode1 development:. Environ. Sci. & Technol., 29(4): 10481058 pp.
Devinny, J. S., M. A. Deshusses, and T.S. Webster (1999). "Biofiltration for Air Pollution
Control", Boca Raton, Florida, Lewis Publishers.
Dullien, F. A. L. (1992). "Porous media: fluid transport and pore structure". 2nd ed, San Diego,
CA, Academic Press, Inc.
Ergas, S. J., E. D. Schroeder, D. P. Y. Chang, and R. L. Morton (1995). "Control of volatile
organic compound emissions using a compost biofilter". Water Environment Research,
67(5): 8 1 6-82 1 pp.

Ergun, S. (1952). "Fluid flow through packed columns". Chemical Engineering Progress, 48(2):
89-94 pp. pp.
Farrugia, V. M. (1999). "The development and properties of biofilms in biofilters". Master of
Applied Science Thesis, Department of Chemical Engineering and Applied Chemistry,
University of Toronto.
Fletcher, M. (1992). "The measurement of bacterid attachment to surfaces in static systems" in
Biofilms-Science.L. F. Melo, T. R. Bott, M. Fletcher and B. Capdeville,
The Netherlands, Kluwer Academic Publishers 603-6 14 pp.
Govind, R., V. Utgikar, W.Zhao, S. Y.,M.Pmatiyar, and D. F. Bishop (1993). "Development
of novel biofilters for treatment of volatile organic compounds (VOCs)". IGT Symposium
on Gas, Oil and Environmental Biotechnology, Colorado Springs, Colorado, November
29-December 3.
Hodge, D. S. and J. S. Devinny (1 995). "Modelingremoval of air contarninants by biofiltration".
Journal of Environmental Engineering, 12 l(1): 2 1 -32 pp.

Effects of Biomass Growth on Pressure Drop in BioJZters

Hodge, D. S., v. F. Medina, Y. Wang, and J. S. Devinny (1992). "Biofiltration: Application for
VOC emission control". 47th Purdue Industrial Waste Conference, West Lafayatte,
Indiana, Lewis Publishers, Inc., 609-6 19.
Holland, F. A. (1973). "Fluid flow for chernical engineers", London, Edward Arnold.
Jain, A. K. (1996). "Section VIII: Pulp bleaching and the environment, Chapter 4: Bleach plant
. .
air emissions" in
C. W. Dence and D. W. Reeve,
Atlanta, Georgia, TAPPI Press 82 1-834 pp.

u
.

Jones, H. R. (1973). "Pollution Control and Chernical Recovery in the Pulp and Paper Industry",
New Jersey, U. S. A., Noyes Data Corporation.
Jones, W. L., R. G. Mirpuri, S. Villaverde, 2. Lewandowski, and A. B. Cunningham (1997).
"The effect of bacterial injury on toluene degradation and respiration rates in vapor phase
bioreactors". Wat. Sci. Tech., 36(1): 85-92 pp.
Kinney, K. A., D.P. Y. Chang, E. D. Schroeder, and K. M. Scow (1996). "Performance of a
directionally-switching biofilter treating toluene contarninated air". 89th Annual Meeting
& Exhibition, Nashville, Tennessee, June 23-28.

Kuwahara, F., Y. Kameyarna, S. Yamashita, and A. Nakayama (1998). "Numerical modeling of


turbulent flow in porous media using a spatially periodic array". Journal of Porous Media,
l(1): 47-55 pp.
Leson, G., J. Rickun, and M. Henson (1991). "Biofiltration- An innovative control technology
for odors and air toxics". 1 991 TAPPI Environmental Conference, 349-354.
Leson, 0.and B. J. Smith (1997). "Petroleum environmental research forum field study on
biofilters for control of volatile hydrocarbons". Journal of Environmental Engineering,
123(6): 556-562 pp.
Leson, G. and A. W. Winer (1991). "Biofiltration: An innovative air pollution control
technology for VOC emissions". J. Air Waste Manage. Assoc., 41(8): 1045- 1054 pp.

Liu, P. K. T., R. L. Gregg, H. K. Sabol, and N. Barkley (1994). "Engineered biofilter for
removing organic contaminants in air". Air & Waste, 44(March): 299-303 pp.
Lowry, O. H.,
N. J. Rosebrough, A. L. Farr, and R.J. Randall(195 1). .J. Biol. Chem., 193: 265275 pp.
Macdonald, 1. F., M. S. El-Sayed, K. Mow, and F. A. L. Dullien (1979). "Flow through porous
media- the Ergun equation revisited". Ind Eng. Chem Fundam., lB(3): 199-208 pp.

Effects of Biomms Growth on Pressure Drop in Biofiters

MacFarlane, S. T. (1998). "Investigation of airflow through compost-based biofilters". Ph. D.


Thesis, Graduate Department of Civil Engineering, University of Toronto.
Madarnba, P. S., R. H.Driscoll, and K. A. Buckle (1994). "Bulk density, porosity and resistance
to aifflow of garlic slices". Dryhg Technology, 12(4): 937-954 pp.
Methods for Soi1 Analysis (1996). "Methods of soi1 analysis". SSSA Book Series, Wisconsin,
Soils Science Society of America, Inc., Madison American Society of Agronomy, Inc.
Mohseni, M. and D. G. Allen (1997). "Biofiltration of a-pinene and its application to the
treatment of pulp and paper air emissions". 1997 TAPPI Environmental Conference,
Minneapolis, Minnesota, May 4-7.
Mohseni, M., G. A. Allen, and K. M. Nichols (1998). "Biofiltration of a-pinene and its
application to the treatrnent of pulp and paper air emissions". TAPPI Journal, 81(8): 20521 1 pp.
Mohseni, M. T. (1998). "Biofiltration of hydrophilic and hydrophobic volatile organic
compounds using wood-based media". Ph. D. Thesis, Graduate Department of Chemical
Engineering and Applied Chemisy, University of Toronto.
Morgenroth, E., E. D. Schreoder, D. P. Y. Chang, and K. M. Scow (1996). "Nutrient limitation
in a compost biofilter degrading hexane".J. Air d; Wate Manage. Assoc., 46: 300-308 pp.
Nielsen, P. H., A. Jahn, and R. Palmgren (1997). "Conceptual mode1 for production and
composition of exopolymers in biofilms". Wat. Sci. Tech., 36(1): 11-19 pp.
Ottengraf, S. P. P., Meestea, J. 5. P., Van Der Oever, A. H. C., and Rozema, H. R. (1986).
"Biological elimination of volatile xenobiotic compounds in biofilters". Bioprocess
Engineering: 6 1-69 pp.
Pedersen, A. R. and E. AMn (1997). "Toluene removal in a biofilm reactor for waste gas
treatment". Wat. Sci. Tech., 36(1): 69-76 pp.
Perry, R. H. and D. Green (1988). "Perry's Chemical Engineerst Handbook". 6th ed, McGrawHill International.
Pinkerton, J. E. (1988). "MACTportion of the Cluster Rule". TAPPIJournal, Sl(2): 99-105 pp.
Sabo, F., U. Motz, and K. Fischer (1993). "Development and testing of hi&-efiiciency
biofilters".86th Annual Meeting & Exhibition, Denver, Colorado, June 13- 18.

Effecfsof Biomms Growfhon Pressure Drop in Bioflliers

Schonduve, P., M. Sha, and A. Friedl(1996). "Infiuence of physiologically relevant parameters


on biomass formation in a trickle-bed bioreactor used for waste gas cleaning". Appt
Microbiol. BiotechnoL , 45: 286-292 pp.
Shareefdeen, 2.and B. C. Baltzis (1994). "Biofiltration of toluene vapor under steady-sate and
transient conditions: theory and experimental results". Chernical Engineering Science,
49(24A): 434704360 pp.
Shareefdeen, Z., B. C. Baltzis, Y. Oh, and R. Bartha (1993). "Biofiltration of methanol vapor".
Biotechnology and Bioengineering, 41(5): 5 12-524 pp.
Silyn-Roberts, G. and G. Lewis (1997). "A technique in confocal laser microscopy for
establishing biofilm coverage and thickness". Wat. Sci. Tech., 36(lO): 1 17- 124 pp.
Singh, A., R. C. Kuhad, V. Sahai, and P. Ghosh (1994). "Evaluation of biomass" in Advances in
EnguirenPe/Rio&&&gy. A. Fiechter, Berlin, Germany, Springer-Verlag.
51: 47-70 pp.
Sinitsyn, V. Y., A. G. Dontsov, and N. A. Gubinov (1991). "Treatrnent of sulfur-containing
gaseous emissions of a kraA mil1 using solid-phase biofilters". Bumazhnaya
Promyshlennost, 6(7): 17 pp.
Smith, F. L., G. A. Sorial, M. T. Suidan, A. W. Breen, and P. Biswas (1996). "Development of
two biomass control strategies for extended, stable operation of highly efficient biofilters
with high toluene loadings". Environ. Sci. Technol., 30(5): 1744- 175 1 pp.
Sorial, G. A., F. L. Smith, M. T. Suidan, A. Pandit, P. Biswas, and R. Brenner (1997).
"Evaluation of trickle bed air biofilter performance for BTEX removal". Journal of
Environmental Engineering, 123(6): 5 30-53 7 pp.
Standefer, S. and C. Van Lith (1993). "Biofiltea minimize VOC emissions". Environmenla1
Protection: 48-59 pp.
Standefer, S. and R. Willingham (1998). "Industrial application of biofilter technologies". 1998
USC-TRG Conference on Biofiltration (an Air Pollution Control Technology), University
of Southern California, Los Angeles, California.
Stewart, P. S., R. Murga, R. Srinivasan, and D. De Beer (1995). "Biofilm structural
heterogeneity visualized by three microscopie methods". Wut. Res., 29(8): 2006-2009pp.
Swanson, W. J. and R. C. Loehr (1997). "Biofiltration: fundamentals, design and operations
principles, and applications". Journal of Environmental Engineering, 123(6): 538-546 pp.
Tate III, R. L. (1995). "Soi1 Microbiology", New York, U. S. A., John Wiley & Sons, Inc.

Effects of Biomass Growth on Pressure Drop in BioJIters

Taylor, S. W. and P. R. JafE (1990). "Biofilm growth and the related changes in the physical
properties of a porous medium 1 . Experimental Investigation". Water Resources Research,
26(9): 2 1 53-2 1 59 pp.
Taylor, S. W., P. C. D. Milly, and P. R. JafE (1990). "Biofilm growth and the related changes in
the physical properties of a porous medium 2. Permeability". Water Resources Research,
26(9): 2 1 6 1-2169 pp.
Taylor, S. W. and J. P. R. (1990). "Biofilm growth and the related changes in the physical
properties of a porous medium". Water resources research, 26(9): 2 1 7 1-21 80 pp.
Togna, P. A. and M. Singh (1994). "Biological vapor-phase treatment using biofilter and
biotrickling filter reactors: practical operating regimes". Environmental Progress, 13(2):
94-97 pp.
Utkin, 1. B., M. M. Yakimov, E. K. Kozlyak, and 1. S. Rogozhin (1989). "Biological air
purification methods". Transiated fiom PriWudnayu Biokhimiya i Mikrobiologiya, 25(6):
723-733 pp.

Van Langenhove, H., E. Wuyts, and N. Schamp (1986). "Elimination of hydrogen sulphide from
odorous air by a wood bark biofilter". Wat. Res.. 20(12): 147 1-1476 pp.
Van Loosdrecht, M. C. M., D. Eikelboom, A. Gjaltema, A. Mulder, L. Tijhuis, and J. J. Heijnen
( 1 995). "Biofilm structuresf'. Wut. Sci. Tech.. 32(8): 35-43 pp.

Vice, K. and R. Carroll (1998). "The Cluster Rule: A summary of Phase 1". TAPPI Journal,
81(2): 91 -98 pp.
Wani, A. H., R. M. R. Branion, and A. K. Lau (1997). "Biofiltration: a promising and cost effective control technology for odors, VOCs and air toxics". J. Environ. Sci. Heafth,
32(7): 2027-2055 pp.
Widerer, P. A. and W. G. Characklis (1989). "Structure and fnction of biofilms". Dahlem
Konferenzen, Germany, S. Bernhard, 5- 1 7.
Williams, T. 0. and F. C. Miller (1992a). "Biofilters and facility operations".
BioCycle(November): 75-79 pp.
Williams, T.0. and F. C. Miller (1W b ) . "Odor control using biofilters". BioCycle: 72-77 pp.

Effects of Biomass G r ~ w t h
on Pressure Drop in Biofilers

Wright, W. F., E. D. Schroeder, and D. P. Y. Chang (1998). "Characterization of regular


transient loading response of a flow-direction-switching vapor-phase biofilter". Air &
Waste Management Association's 9 1st Annual Meeting & Exhibition, San Diego,
California, June 14-18, paper 98-WAA.07.
Wright, W. F., E. D. Schroeder, D. P. Y. Chang, and K. Romstad (1997). "Performance of a
pilot-scale compost biofilter treating gasoline vapor". Journal of Environmental
Engineering, 123(6): 547-55 5 pp.

Zhang, X., P. L. Bishop, and M. J. Kupferle (1997). "Measurement of polysaccharides and


proteins in biofilm extracellular polymers". Second International Conference on
Microorganisms in Activated Sludge and Biofilm Processes, Berkeley, Califomia. July
2 l-23,55 1-554.
Zich, T., F. Proll, and A. Fnedl(L996). "Pellets, a new type of filter material for biofiltration".
Conference on Biofiltration ( an Air pollution Control Technology), Univesity of Southem
Califomia, Los Angeles, October 24 & 25, 128- 133.

Effects of Biotnuss Growth un Pressure Drup in Biojlters

A- l

Appendix A

APPENDlX A:
Biofilters

Calculation of Nutrient Requirements for 60th

Macronutrient requirements were calculated using a C/N ration of 30: 1. The fertilizer pellets
were assurned to be the only supply of macronutrients. The calculation considered 100 %
removal, based on maximum removal capacities observed in biofilters in other experiments
conducted in this research group (Mohseni, 1998). Macronueients were supplied in excess
since the removal capacities of the biofilters were lower than the maximum values assumed
here.

Bed specitkations:
Bed height/ section: 0.30 m
Bed diamter: 0.28 m
Bed volume/section = n*(0.28m/2)~*30m= 0.0185 m3
Total be volumehiofilter = 4* 0.0185m3= 0.074 m3

Loading rates:
Operating tirne:
Total loading

Molecular weight
Density (liquid)
Number of carbons
Carbon loading

Methanol

a-pinene

150 g methanol/m3bed/h
120 days
2880 h
150*2880*0.074 =
= 31921 g methanol
34.04 g/mol
0.79 14 g/mL
1 (Cho)
3 1921/(34.04* 12.01) =
= 11965 g C

40 g agpinene/m3bedfh
120 days
2880 h
40*2880* ,074 =
= 85 12 g a-pinene
136.24 g/mol
0.858 g/mL
10 (ci0Hi6)
85 l2/(136.24* 10*12.01) =
= 7504 g C

Total carbon/biofilter
g nitrogenhiofilter
Fertiliser's nitrogen content
Fertiliser requiredhiofilter

Fertiliser requiredhection

Fertiliser required/m3 bed

11965 + 7504 = 19469 g C


19469/30 = 649 g nitrogen
20 % (refer to Table 3.2)
649/0.20 = 3245 g fertiliser
3245/4 = 8 11 = 800 g fetiliser/section
32490.074 = 4385 1g fertiliser/m3bed
= 44 kglm3bed

Effects of Biomass Growth on Pressure Drop in Biofilrers

APPENDIX B: Calibration Data for


Chromatograph

the Varian

3600CX

Gas

The GC was calibrated based on air samples containing a known arnount of moles of methanol
or a-pinene, which were prepared in sealed Teflon bags. The following sections show the
procedure and calibration data for each compound.

B.1

Calibration for methanol at a constant oven temperature of 120 O C .

Calibration for GC (methanol)


Methanol

M.W.
density

M.P.
B.P.

BOTTLE #

32.04 glmol
0.7915 g/mL
-97.7 C
64.7 C

Added
Added
Added
Air
methanol methanol
(PL)
mols
(mL)

ppmv-aw*
Moles of
(based on
methanol in
250
of air 250
air)

Sample prepared fkom sample M#O


Sample prepared fiom sample M#O

1800
1800
1800
1800
1800
1800
1800
1800

M #O

900

0.2
0.5
1.5
3
5
6.5
8
1O

4.94E-06
1.24E-05
3.71 E-05
7.41 E-05
1.24E-04
1.61 E-04
1.98E-04
2.47E-04

10
2.47E-O4
6.86E-08
*Based on molar volume = 24.5 Lhol air (P=l atm; T=298 K)

6724.8

Effects of Biomuss Growth on Pressure Drop in Biofilers

Calculation sample
Total moles methanol in sample (eg. 0.2 MLadded)
=V*pIM.W.
=(0.2 PL 0.791 5 glmL 1 32.04 gimol) 1 1000 PLImL
= 4.94E-6 mol

hiloles methanol in 230 PL of gas (eg. 0.2 PL methanol in 1800


mL air)
=mol methanol tot / V*S.V.
= (4.94E-6 mol / 1800 mL)+250pL I i O O O pL/mL
= 6.8BE-10 mol (in 250 PL sample)

ppm methanol in air (eg. 6.86E-10 PL methanol in 250 PL sample)


=(mol methanol inj.1V samp.) a M.V. 1E6 mol air
=(6.86E-10 mol 1 250 PL) a 24.5 L /mol air a 1 E6 *l
E+6 pLlL
= 67.2 ppm (air)

GC calibration cuwe for mthanol

500

1000

1500

2000

2500

3000

3500

4000

ppmv methanol

Eflects of Biomass Growth on Pressure Drop in ~ i o f i l t e k

Experimental data:
Peak area (counts)
Concentration
Average of 3 runs Standard deviation
ppmv methanol

Regression equation:
Peak area = Slope x concentration + intercept

Parameter

R~
Slope
Intercept
Set intercept to zero

Estimate
0.999
392
2703

Confidence interval
Lower 95 % Upper 95 %
389
-4756

395
10162

R~
Slope

B.2

Calibration for a-pinene at a constant oven temperature of 120 O C .

Calibration for GC tn-~inene)

M.W.
density

M.P.
B.P.

Effects ofBiomass Growth on Pressure Drop in Bioflters

Appendix B

B-4

e
e
5
ppmv-air

BOTTLE #

methanol in
(mL)
(+)
moLp
250
of air
C# 1
1800
0.1
6.30E-07
8.75E-11
C #2
1800
O .3
1.89E-06
2.62E-10
C #3
1800
O .5
3.15E-06
4.37E-10
C #4
1800
0.8
5.04E-06
7.00E- 10
C #5
1800
1.2
7.56E-06
1.OSE009
C #6
1800
1.8
1.13E45
1S7E-09
C #7
1800
3
1.89E-05
2.62E-09
C #8
800
2
1.26E-05
3.94E-09
C #18
1800
2.5
1S7E-05
2.19E-09
C #19
1800
4
2.52E-05
3SOE-09
* B a d on molar volume = 24.5 Ymol air (P= 1atm; T=298K)
Air

methanol

methanol

Calculation sample
lotal moles a-pinene in sample (eg. 1.2 PL added)
=V P 1 M.W.
=(1.2 PL * 0.858 g/mL / 136.24 glmol) / 1000 vL/mL
= 7.56E-6 mol
.

-. ..

(based on
250 ,& air)

8.6
25.7
42.9
68.6
102.9
154.3
257.2
385.7
2 14.3
342.9

CALIB
LEVEL
1
2
3
4
5
6
7
8
9
10

-..

Moles a-pinene in 250 PL sample (eg. 1 . 2 ~ Lin 1800 mL


sample)
=mol a-pinene tot / V9S.V.
= (7.56E-6 mol / 1800 mL)'250pL / 1000 ~LlrnL
= 1.05E-9 mol (for 250 PL sample of air)

ppm a-pinene in air (eg. 1.05E-9 mol in 250 PL sample of air)


=(mol a-pinene inj.1V samp.) ' M.V. 1E6 mol air
=(1.05E-9mol / 250 PL) 24.5 L /mol air 1E6 1' E+6 PLIL
= 102.9 ppm (air)

Effects of Biomass Growth on Pressure Drop in Bio#lters

OC caiibration cuwe for, -pinene

Experimental data:
Concentration Peak area (counts)
ppmv ,-pinene
Average of 3 runs
9
142546

Standard deviation
7962

Regression epuation:
Peak area = Slope x concentration + intercept
Parameter

Estimate

Confidence intewal
Lower 95 % Upper 95 %

R2
Slope

Effects of Bioonuzss Growth on Pressure Drop in Biofiters

APPENDM C: Biomass Quantification Profiles for the Nova Inert and


Wood Cbip Biofilter as COD, TS, and VS during the biofiltration of methanol
I

I
I
I

E
a

100

I
I
I
I

I
I

O. 1

0.2

0.3

0.4

0.5

I
I
I

I
I
I

1-

- - . - - -Day 56
4D ~ 84
Y
-Day
111

0.6

0.7

0.8

**

Day 148

0.9

1.1

Bed depth (m)

Figure C.l
Biomass concentration profiles as Chemicai Oxygen Demand (COD) along the Nova
Inert biofilter measured in 4 different days during the biofiltration of methanol

-Day

49

+Day 76
D

y IO5

+Day 148
---

Figure C.2
Biomass concentration profiles as Chemicai Oxygen Demand (COD)along the wood
chip biofilter in 4 different days during the biofiltration of methanol

Effects of Biomaps Growth on Pressure Drop in Bioflters

Figure C.3
Biomass concentrationprofiles as Total Solids (TS)dong the Nova Inert biofilter in 4
different days during the biofiltration of methanol

Figure C.4
Biomass concentration profiles as Total Volatile Solids (TVS) dong the Nova hert
biofilter at 4 bed sampling days

Eflects of Bioniars Growth on Pressure Drop in Bioflters

Appendix C

C-3

Figure CS
Biomass concentration profiles as Total Solids (TS) along the wood chip biofilter in 4
different days during the biofiltration of methanol

Figure C.6 Biomass concentration profiles as Total Volatile Solids (TVS) dong the wood chip
biofilter in 4 different days during the biofiltration of methanol
--

Eflects of Biomnss Growth on Pressure Drop in ~iofilers

APPENDIX D: Comparison between Biomass Measurements as


CODTTS and Detached Biomass in the Wood Chip Biofilter
--

oDay49
Day76 I
~ D a 105
y
x Day 148
1

10

15

20

g Mdayerlg dry substratum

Figure D.l
Comparison between biomass measurements as mg COD/g dry substratum and as g
biolayerlg dry substratum in the wood chip biofilter during the biofiltration of methanol

10

15

20

g bidayerlg dry substratum

Figure D.2 Comparison between biomass measurements as mg TS/g dry substratum and as g
biolayedg dry substratum in the wood chip biofilter during the biofiltration of methanol
Effects of Biomass Growth on Pressure Drop in Biofilters

Appendix E

E-l

APPENDIX E: Transient Conditions after Bed Sampling and


Remixing, before Switching over to a-pinene
I

I
I

SBdjal 2

i
I

3 ~ - ! % d y s t ~ e e t k h Fenixing (day 146)


+2 h after restarting

1
1
I
I
I
I

section 4

Figure E.1
Methanol axial concentrations along the Nova Inert biofilter after sampling and remixing the
bed, just before switching over to a-pinene. A discontinuous line indicates the approxirnate location of each
section

Figure E.2
Methanol axial concentrations along the wood chip biofilter, after sampling and remixing
the bed, just before switching over to a-pinene. A discontinuous line indicates the approximate location of
each section

Efects of Biomass G r ~ onhPressure Drop in Bioflters

APPENDIX F: Transient Conditions after Switching over from


Methanol to a-pinene
30
I

sedion 1

I
I

Section 2

I
I

section 3

I
I
I

section 4

25
Q
Y

c
0
20

E
C
g

CI

15

g IO
Q)
9-

35
O
O

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.1

Bed depth (m)

Figure F.1
Pinene axial concentrations along the Nova Inert biofilter afier switching over from
methanol to a-pinene.

O. 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

ed depth (m)

Figure F.2
Pinene axial concentrations along the wood chip biofilter after switching over from
methanol to a-pinene

Eflecfs of Bionas Gtowfhon Pressure Drop in Biofiters

Appendix G

G-1

APPENDIX G: Pressure Drop Profiles during the Biofiltration of a-pinene


I
I
I
I
I
I

sedion 1

I
I
I

section 2

A
w

l
I

..

I
I

section 3

"

I
l
I

section 4

I
I
I
I

Fl

I
1
I
1 ,

v.-.- - -- ...

1
I
I
I
I

1
I
1
I

I
I

0.1

0.2

0.4

0.3

0.5

0.7

0.6

-w- - -

-.

+Day 150 (start penod 5)


+Day 181 (end period 5)

+Day 183 (start penod 6 )


+Day 209 (end penod
6)
.. .
A

-- -

0.8

--

. . .

0.9

1.1

Bed depth (m)

Figure G. 1 Cumulative pressure drop profiles dong the Nova Inert biofilter at the beginning
and at the end of each undisturbed-bed penod of operation during the biofiltration of a-pinene. A
discontinuous line indicates the approximate location of each section
1
1

+Day 150 (start peflod 5)t

section 3

section 4

I
I

+Day 181 (end period 5)


Day 183 (start period 6)1
+Day 209 (end period 6 )

O. 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Bed depth (m)

Figure G.2
Cumulative pressure drop profiles dong the wood chip biofilter at the beginning and
at the end of each undisturbed-bed period of operation during the biofiltration of a-pinene. A
discontinuous line indicates the approximate location of each section

Effects of Biomuss Growth on Pressure Drop in Biofiters

APPENDIX H: Biomass Quantification during the Biofiltration of a-pinene

I
O

O. 1

+~

a216
ywood chips
--.- - -

0.2

0.3

0.4

0.5

0.6

O.7

O.8

O.9

Bed depth (m)

Figure H.1 Profiles of biomass quantification as Chemicd Oxygen Demand (COD) along the Nova
Inert and wood chip biofilters at each bed sampling day during the biofiltration of a-pinene

Figure H.2
Profiles of biomass quantificationas Total Solids (TS)dong the Nova Inert and wood
chip biofilter at each bed sampling day during the biofiltration of a-pinene

Effects of Biornass Growrh on Pressure Drop in Bioflters

182 Nova lnert

-&y

+Day 216 - NOM lnert


-ay
-Day

l82-Miod chips
216 Hlood chips

Figure H.3
Profiles of biomass quantification as Total Volatile Solids (TVS)along the Nova
Inert and wood chip biofilters at each bed sampling day during the biofiltration of a-pinene

- -...- .- - - - --- . -

10

8 -

W a n2

- --- -- -

6 -

0.1

. .

d o n4

m o n1

Day 182 Nova lnert


'+Day
216 Nova lnert
Day 182 wood chips
+Day 2 16 - wood chips

m o n3

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Bed depth (m)

Figure H.4
Profiles of biomass quantification as total protein dong the Nova Inert and wood
chip biofilters at each bed sarnpling day during the biofiltration of a-pinene

Effects of Biomass Growth on Pressure Drop in Biofiters

Estimation of the Yield Coefficient for the Nova lnert


APPENDlX 1:
and Wood Chip Biofllter
Figures 1.1 and 1.2 show the linear regression for TVS as a function of methanol removed for the
Nova Inert and wood chip biofilter, respectively.
200

Cumulative removal (g methanol)

Figure 1.1

Estimation of the yield coefficient for the Nova Inert biofilter

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Cumulative removal (g methanol)

Figure 1.2

Estimation of the yield coefficient for the wood chip biofilter

EflPects of Biomass Growth on Pressure Drop in Biofilers

Appendix J

APPENDIX J:
Biofilters

J- 1

Calculation of Reynolds Numbers (N&

for Both

J. 1 Calculation
Reynolds nurnbers were calculated for each measurernent of pressure drop as a function of air
flow rate in each biofilter. The expression used is (Holland, 1973):

where p is the air density, p is the air dynamic viscosity, u is the air superficial velocity, E is bed
porosity, and so is the surface area per unit volume of solid matenal. Equation J.1 assumes that
the equivalent diameter (de) for flow through the packed bed can be defined as 4 times the cross
sectional flow area divided by the flow perimeter, as show by the following equation for a unit
height of bed;

The surface area per unit volume of the solid material in the bed (sn) was estimated for nonspherical paxticles assuming an average particle diameter, d,,:

(Eq. 5.3)
where v is a shape factor equal to the quotient of the area of a sphere equivalent to the volume
of the particle divided by the actual surface of the particle (y = 1 for spheres).

Parameters:
1. Air dynamic viscosity at 40 O C near ambient pressure: 1.85~10'~
Pa.s (Peny and Green,
1988)
2. Air density by Ideal Gas law:
rnolHzO/mol air):
Air at T = 3 13 K, P = 1 atm, 100%relative humidity (y,,,,=0.0726
Average molecular weight of humidified air = 0.9274(29g/mol)+0.0726(18g/mol)
= 28.2 g/mol
(P)MW - (101300~ax28.2~
/ mol)
lkg
= 1.089kg / m3
P=
RT
( 8 . 3 ) l O O O g
3. Air superficial velocity (u):
Cross-sectional area of the bed = n ( 0 . 2 8 m ~ =
) ~6.158x10-2 m2
-2 2
u=(Qrn3/~)/6.158~10
m
4. Packing materials:
Particle size distribution: refer to Table 3.1 for both materials

Effects of Bionurs Growth on Pressure Drop in Biofiters

Mean particle diameter estirnated based on screening size and volumetnc fraction of size
in mixture:
d, = L ((average size)(volurnetric fraction))
Nova Inert pellets:
d,, = [(l .l+.7)/2](0.33) + [(.7+.5)12](0.67) = 0.7 cm = 0.007 m
y = 0.85 (W = 0.75 for sand of various types; y = 0.83 for rounded sand)
Wood chips:
dp= [(3S+2.3)/2](0.07)+[(2.3+l.6)/2](0.33)+[(1.6+l. 1)/2](0.40)+
[(1.1+0.7)/2](0.20) = 1.6 cm = 0.016 m
w = 0.3 (y = 0.3for Raschig rings; = 0.28 for mica flakes; y~ = 0.69 for cork)
S. Bed porosity
Estimated fiom the average wet porosity measurements for the four sections of the
biofilter
Nova Inert
Wood chip
First penod
E = 0.48
E = 0.48
Second period
E = 0.44
E = 0.41
Third period
E = 0.39
E = 0.39
Fourth period
E = 0.38
E = 0.34

Effects of Biomass Growth on Pressure Drop in BioJTIlers

Appendix J
5.2

5-3

Data

NOVA INERT BlOFllTER


REYNOLDS NUMEERS

1
op-no

Ur liaw rate (Umir.)

dw

~ o 2 10.y
.d - 57 a31
ir fiaw rats (Umin)
Total AP ( P d m bed)

1281

751

I
(NRJb
14

153
136

4:

87
126

131
153
97
83

4(

lt7
133
1 53
87

5:

1 27
1 37
151
113
Ba
81

125
140

82

151

1l a
93
74

mnoaJ(uq

Lir ilow rate (Umin)

Operaling day
9

FWOd 4 (Oiy 112 -1M )


rlr flaw rate (Umin)
Total AP ( W m bed)

1 O(

128l

1241

(NRJb
13

10'

Ili

Table J. 1
Estimation
Reynolds
measurement
function of air flow rate for the Nova Inert biofilter during the biofiltration of methanol

Effects of Biomass Growth on Pressure Drop in Bioflers

Avvendix J

5-4

REYNOLDS NUMBERS

I
Ur fiow rate (Umin)

1
2
6
133
114
98
81
126
129
114
99

79
65
38
127
113
96
75
61
43
128
il5
98
82
65
45
120
111

94
82
62
40
128
il3
95
70
60
42
127
1O7
92
75
57

37

124
i la
91
76
59
41

Table 5.2
Estimation of Reynolds numbers for each pressure drop measurement as a function
of air fiow rate for the wood chip biofilter during the biofiltration of methanol

Effects of Biomass Growth on Pressure Drop in Biofilrers

p.iiw 3 (oay

Iperating day

lAir ffowrate (Umin)


831

r7 104)

Totai AP (Palm b a i )

124

Table 5.2 (contd.)


Estimation of Reynolds
pressure drop measurement
function of air flow rate for the wood chip biofilter during the biofiltration of methanol

Effectsof Biomuss Growth on Pressure Drop in Bioflters

APPENDIX K:

Pulse Tracer Tests

K.1 Main results and discussion


The unimodal curves obtained fkom the tracer tests showed that there was not any apparent
channelling through either biofilter at the time of the experiments. The average tracer curves
obtained in the middle of the fourth undisturbed-bed penod, on day 126, during the biofiltration
of methanol, are presented in Figure K.1. Figure K.2 shows the average tracer curves before
and after remixing the bed, and Figure K.3,the average tracer curves from day 167 during the
biofiltration of a-pinene. The earlier peak in the Nova Inert biofilter pulse response curve from
day 126, and the wider curve for the wood chip biofilter suggest that higher pomsity was
available in the wood chip biofilter than in the Nova Inert biofilter, and that there was more
dispersion in the wood chip biofilter than in the Nova Inert biofilter. An increase in bed
porosity would be shown by the displacement of the unimodal curve towards the left, indicating
a higher retention time. Even though more dispersion in the wood chip bed could have been
expected because of the variation in shape and size of the chips, a higher porosity in the wood
chip bed was not likely to occur since high biomass accumulation and compaction took place in
this biofilter. This is also inconsistent with measurements of wet porosity (section 4.1.4).

Figure K. 1 Nomalised response curves to methane pulse injections coaducted at the inlet of
the Nova Inert and wood chip biofilters on day 126 during the biofiltration of methanol. C
stands for tracer concentration and Cmax for the maximum tracer concentration measured at the
outlet of the biofilters.

Effectsof Biornass G m t h on Pressure Drop in Biofilers

Appendix K

K-2

Figure K.2
Normalised response curves to methane pulse injections conducted at the inlet of
the Nova Inert and wood chip biofilters on days 147 and 148, before and after remixing previous
to switch over to a-pinene. C stands for tracer concentration and Cmax for the maximum tracer
concentration measured at the outlet of the biofilters.

Figure K.3 Normalised response curves to methane pulse injections conducted at the iniet of
the Nova Inert and wood chip biofilters on day 167 during the biofiltration of a-pinene. C
stands for tracer concentration and Cmax for the maximum tracer concentration measured at the
outlet of the biofilters.

Effects of Biomass Growth on Pressure Drop in Bioflters

Appendix K

K-3

The effects of remixing the beds on the flow characteristics in both biofilters are reflected in
Figure K.2. The tracer c w e s before and after remixing the Nova Inert bed (Figure K.2) show
that there seemed to be no particular effect of remixing on the airfiow. On the other hand, for
the case of the wood chip biofilter since &er remixing the maximum of the curve was reached
faster, this would suggest that the porosity in the bed was reduced. The observations made fiom
these curves seem to oppose what would be expected to happen, that is, an increase in the
porosity of the bed after remixing due to bed expansion. It should be pointed out that new wood
chips needed to be added to the bed d e r remixing (7% in volume more). This certainly
changed the characteristics of bed, and may not allow for cornparisons of the response curves
before and after remixing.
nie response curves fiom day 167 during the biofiltration of a-pinene indicate that both
biofilters had similar flow characteristics, which was not expected because both biofilters had
different bed porosities, different biomass concentrations, and different bed volumes, after the
wood chip bed had experienced compaction.
Tracer tests are employed as a means of obtaining information about the flow characteristics in a
reactor, and have been used in biotrickling filtea (Karamanev et al., 1994) and biofilters
(Kiared et al., 1996). For the case of a packed reactor, such as a biofilter, in addition to
describing the air movement through the biofilter, tracer tests can provide information to
calculate the effective interparticle porosity of the bed (Deshusses et ai., 1995). This is
advantageous for effective porosity estimation since it does not require the disturbance of the
bed.
In this study, tracer tests were conducted in the biofilters in the middle of the fourth operation
period, before and after switching over to a-pinene, and during the first period operating with apinene. Even though these expenments would have been useful during the other undisturbedbed periods of operation with methanol, they were not carried out since the set-up was not
available at that time. Unfortunately, the results obtained f h m the tracer tests that were
conducted did not provide reliable information about the flow characteristics through the
biofilters.
Even though the unimodal shape of the tracer tests indicates that no major air channelling was
taking place in the biofilters, it is believed that the actuai tracer curve for the biofilters could
have been masked by the operation of the total hydrocarbon analyser (THA) used to measure the
methane (tracer) concentration at the outlet of the biofilters. The analyser pumped air from the
outlet of the biofilters and passed this sample through the flame ionisation detector. A
description of the flow through the THA only was obtained with pulse tracer tests conducted
using just the line fiom the outlet of the biofilters and the hydrocarbon analyser. The tracer
c w e fiom these tests showed a unimodal distribution of the concentration. In addition, there
was more fiow dispersion in the analyser than expected, which deviated the flow fiom plug
flow.
Further analysis of the airfiow through the THA was achieved by using the tanks-in series (TIS)
and the dispersed plug flow (DPF)models. These models for non-ideal flow characterise the
fluid flow through a vesse1 using one adjustable parameter: the number of completely stirred
tanks in series (N) in the TIS model, and the dimensiodess quotient of convective flux over
dispersed flux referred as the Peclet number (PeL)in the DPF model. As calculated in Appendix
L, the response time curve fiom the analyser alone gave an N value amund 11 and a PeL
between 21 and 25. Considering that for the TIS model N=l corresponds to an ideal mixed flow
and for N>30 plug flow is virtually achieved, and given that for the DPF model PeL+O
approximates to mixed flow and PeL+ 00 approximates to plug flow, it was concluded that the

Effects of Biomass Growth on Pressure Drop in Biofilers

Appendix K

K-4

flow in the analyser alone was subject to mixing, which could have made indistinct the flow
characteristics through the biofilters.
Moreover, a similar analysis using the parameter N for the tracer curves obtained from tests
where the biofilters were comected to the THA showed that N was higher than 1 1, tending the
flow more to plug flow for these cases (See Appendix L). This proved the inconsistency of the
tracer tests since the flow of the tracer through the THA could not have changed the
characteristics of the flow through the biofilters to be closer to those of plug flow. In any case,
it would have been expected that the THA would have controlled the response fiom the tracer
pulses, so that N would have been lower than 1 1, indicating strong flow mixing. There also
exists the possibility that the THA did not affect the tracer response. Nevertheless, the
porosities estimated fiom the tracer tests varied fiom 0.75 to higher than 1, which were very
high porosities or senseless, even though time corrections were considered due to the THA time
delays. The high values of porosities could also have k e n due to a high airflow rate (or
superficiai velocity) used for the tracer tests, which was constant and the sarne for al1 the
experiments, around 0.02 m3/s (120 L/min). MacFarlane (1998) found that the effective
porosity detennined using tracer curves increased when the superficial gas velocity increased
when conducting tracer tests in a compost-wood chip bed. Macfarlane explained these changes
based on the dependence of the mean residence time on the coefficient of dispersion and linear
velocity.
Consequently, a quantitative analysis of the tracer response curves did not provide useful
information about effective porosities in the bed, or even effects on flow before and afier mixing
the bed due to the probable interference from the THA operation. However, a qualitative
analysis of the response curves could indicate that no major channelling occurred in the
biofilters at the time of the tests. This was probably the case, because it is believed that the
THA could not have transformed the response cuves so significantly as to change the response
fiom a bimodal cuve (major channelling) to a unimodal (no major channelling) curve. Some
minor channelling could have happened, although not able to modify significantly the tracer
response curves. Mysliwiec et al. (1996) studied channelling in biofilters, and concluded that a
biofilter may operate for some time with a channel in formation without displaying symptoms of
channelling until the channellised flow reaches a cntical point. Also based on their results, it
c m be concluded that the greater the number of channels, the higher is the probability of the
tracer curve of being unimodal. MacFarlane (1998) obtained unimodal tracer curves fiom pulse
tests with a compost-wood chip bed, even though channelling was present within the bed. By
sampling a tracer at the same bed height but different locations in the compost-wood chip bed,
MacFarlane deduced the presence of channels because of the different residence times obtained
at each location, based on the tracer curves.
In this study, some variations in radial concentrations of a-pinene were measured in the Nova
Inert biofilter at the end of the experimental run, which could indicate this type of channelling.
Based on this expenence, it is recommended that tracer tests be conducted in such a way that
any interference fiom the equipment quantifying the tracer concentration be able to be
accounted for. This could involve tracer step changes, instead of tracer pulses, and the
measurement of the tracer concentration at the inlet and outlet of the biofilter. In addition,
smoke tests and radial measurements of the pollutant concentration at different bed Ievels could
help to gain insight into the characteristics of flow through the biofilters.

Effects of Biomass Growth on Pressure Drop in Biofiters

K.2

Residence Time Distribution (RTD) Function Analysisl Calculation Sample

Table K.1 presents the calculation of the first moment (mean retention time) and second
moment (variance) about the origin for the RTD hction obtained fiom the first pulse tracer test
conducted in the Nova Inert biofilter on day 147. A methane mass balance was also conducted.
Similar analyses and calculations were conducted with the rest of the tracer data fiom other
tests.
MFILtER I (NOVUNtRt)
15.Apr-08
T

m mahane 98%
[ GaugsP(bar)

Run 1

0.B

1 AbsolutsP(am)1

T(C dogme)

1.O9

25.7

(K)

2g0.85

[ V injsded (mL) 1 CH, pulse moks (No)


i
8
1
6.10t-04

Mean tempeiatum bknier ( C degree)


Wet air riowr rate Qo (Umin)

1atrnUonTirna MItrlbuUon(RTD) funttlon anrlyrk


listogram M h o d Ibr calcuiaiing a m under the ~ M ! I
ppmv mhane

0.44
0.32
0.78
0.44
0.42
0.64
0.64
0.64
0.50
0.62
0.48
0.44
0.44
0.82
0.71
0.71
0.68
0.64
o. 64
0.53
0.w
0.64
0.43
0.68
0.64
0.68
0.42
0.53
0.64

E(t)
(1JJ)
7.55t-05

5.53E-05
t.30E-04
7.55E-05
7.14E-05
l.lOE94
1.10E-04
1.10E-04
9.B6E-05
1.06E-04
8.35E-05
7.55E-05
7.55E-05
1.O6604
1.22E-04
1.22E-04
1.16E-04
1.10E-04
l.10E-04
9.16E-05
1.48E-04
1.10E-04
7.14E-05
1.16E-04
1.1OE-04
1.16E-04
7.14E-05
9.16E-05
1.l
OE-04

t t ( 1 ) At
S

O.O0k+dO
5.53E-O5
2.60E-04
2.27E-04
2.86E-04
5.48E-04
6.58E-04
7.68E-04
7.97E-04
9.51E44
8.35E-04
8.31E-04
9.06E-04
1.37E-03
1.70E-03
1.83E-03
1.85E-03
1.ME-03
1.97E-03
1.74E-03
2.9BE-03
2.30E-03
1.57E-03
2.66E-03
2.63E-03
2.89E-03
1.86E-03
2.47E-03
3.07503

'able K.1
Retention Time Distribution function analysis for t racer tests conducted in t
Nova Inert biofilter on day 147 (run 1)

Effects of Biomass Growth on Pressure Brop in Bioflters

Appendix K

K-6
1'

W)At
s2
0.09
O 10
0.07
O 11
0.08
0 12
0.09
O 14
0.24
0.25
0.43
O 53
0.73
1.O7
1.57
2.18
2.98
3.97
5 26
6.90
8.75
11.01
11.51
14.08
17 O4
23.30
27 43
31.44
36 11
41.08
46 10
51 22
56.60
62.12
67 47
72.80
78.69
83 28
89.08
94 22
98.43
103 19
106 81
110 75
114.88
118.19
121.O8
123.40
125.72
126.97
129.25
130.24
130.03
130.77
t 30.65
130.32
129.53
129.32
128.28
126.76
125.07
123.64
121.75

*ableK. 1 (contd.) Retention T h e Distribution funcl


Nova Inert biofilter on day 147 (run 1)

analysis for tracer

n the

Eflts of Biomass Growth on Pressure Drop in Biofiters

Tme
No

ppmv mehme

t (0)
93
94
95
96

93
94
95

97

98

98
99
100
101
102
103
104
1OS
106
107
108
109
110

97
98
99
100
101
102
103

For t, =

104
105
106
107
1oa
108
Il a

For nmalizing C(t) to E(t) use (='At

+ C(tT)la) =

0.000228

m
m(ml
--

---

able K.1 (contd.) Retention Tirne Distribution fiinction analysis for tracer tests conducted in
the Nova Inert biofilter on day 147 (nui 1)
Efects of Biomass Growth on Pressure Drop in Biofiters

mol CM
c(ma)

1.95Em

9.70E-00
9.79E4
9.79E-08
8.979(1
7.91 4 1
7.59E-08
ll.03E4
7.91 4 8
7.31E98
6.06E-M
6.86E-08
6.UE-00
S.89E-00
4.ME-01
5.52E-00
4.04E-00
4.OBE-08
4.24E-08
4.56EQn
4.43E-00
3.97E-(M
3.51E-00
3.69E01
3.37E-08
3.37E-08
2.96E-08

156
157
1Sll
158
100
161
162
163
164
165
186
167
180
169
170
171
172
173
174
f 75
176
177
178

Lmght (m):

cwnaer (in):
0.67 kw me (Umm ):
C#racted fml mlanlkn Ilme
:omcWd main nuntlon tlme (THA nrponw Umm)
THArwponn~

bn-

Cornctrd morn ntontlonUme


Iiterlrl bilrnce
Ids of methme hjedd

tm

No=

6.10E-04

82.58 s

24.3 s

QI

moi malhwie

Mdr d mahanr meuured bued ai rsrpmss mm


No=
4.66E-04
m d mahane

Vokma lnvohrsd h Iricrdudy W (Lp

~np.tkld
v d u h~ b
b=
VU (L) =

119.24

~ ~t ~ ~b

4t

T U volume d bkliner ured in tram les1


diametw (ml
0.28
Total h d ~ h(m)
t
1.65
~0kime~t BI(L)
102.42

'able K. 1 (contd.) Retention Tirne Distribution fiinction analysis for tracer tests conducted in the
Nova Inert biofilter on day 147 (run 1)
K.3

Result from tracer tests

The results fiom the tracer test are s h o w in table K.2 below.

Effects of Biornas Gmwh on Pressure Drop in Bioflers

'Tracer tnt parrmod on Harth 25


(Day 126)

~ o v lnert
a ontbr

'Tracer trot prrormaa on mil116, '13911


(Day 148 a h r nmixlng)
RUn
Mein nwntion urne a (0)
(a2)
Vadame ot th.diotribution,
ai (8)
Comctod marn rebntlon Ume (a)
MOi8W InjOCWd (niOb)
Methano qmnfhd b r w d on moponw curw (mok)
Percent ermr
Total voluma ot bod lnclucnng vord spic^, Vb (L)
Volume calculi(.d from tracer otudy, Vtr (L)
Unpaclred volume In blditor, Vt-Vb (L)
Void vdume in k d (L)
Elhctlw k d poraity

Wood Chip Bionlter

Nova Imrt Biofilbrr

Wood Chip Biofilter


3

82 1

831

84

911

!d

uun
Mern retentionUme tm (O)
(aa)
Varirme of th.diotribution,
al (0)
Comcted moan rotonuon timo (s)
M ~ V ~ I rnjactoa
U
(mol.)
Methrrn qwntifled bmod on nrpome cunn (mok)
Parcant a m r

Vduma cilculrtmd from tmcor atudy, Vtr (L)


Unprckodwlume in biotilbr, Vt-Vb (L)
Void Wume in k d (L)
EWctiva k d porority

~ o v ra ~ rWofiltlr
t

uun
Mein rebntlaci Ume tm (O)
Variance at the dlrtilbution, (sa)
et@)
Comced maan ratontionUme (s)
mamm inJoctM[moro)
Mottirna qwntifiod bsod on mponw c u m (mok)
Parcont error
T ~ volume
I
ot ma InClUdInavoia apmce, vb
Vdume calcukbd tram tracer otudy, Vtr (L)
Unpackad vai~rnein biollbr, Vt-Vb (L)
Vdd volume In k d (L)
Elncuvo k d poraity
-

wooti chip ~ ~ o n ~ t o r
3

:,,

rable K.2

50
0.8 1

49

50

0.79

0.80

48
0.84

47
0.82

46
0.80

Results fio& the aaalysi of the response tirne distribution function obtained

fiom tracer snidies performed in the Nova Inert and wood chip biofilter in 3 different days

Efects of Biomass Growth on Pressure Drop in Biofiters

Appendix K
K.4

Tracer curve for the THA

Figure K.4
K.5

K-10

Response curves obtained fiom tracer tests perfonned in the THA

References

Deshusses, M. A., G. Hamer, and 1. J. Dunn (1995). "Behavior of biofilters for waste air
biotreatment. 2. Experimental evaluation of a dynarnic model". Environ. Sci. & Technol..
29(4): 1048- 1058 pp.
Karamanev, D. Ga, Belanger, MX., Chavarie, C., Chaouki, J., and Mayer, R. (1994).
"Hydrodynarnic characteristics of a tncking bed of peat moss used for biofiltration of
wastewater". The Canadian Journal of Chemiccll Engineering, 72:4 1 1-4 17 pp.
Kiared, K.,Bibeau, L., Bnezinski, R., Viel, G., and Heitz, M. (1996). Biological elimination of
VOCs in biofilter". Environmental Progress, 15(3): 148- 152 pp.

MacFarlane, S. T. (1998). "Investigation of airfiow through compost-based biofilters". Ph. D.,


Graduate Department of Civil Engineering.
Mysliwiec, M. J., Schroeder, E. D., and Kawas, M. L. (1996). "Flow characterization of
biofilters". 89th h u a 1 Meeting & Exhibition, Nashville, Te~essee,June 23-28.

Effects of Biomass Growth on Pressure Drop Ni Biojilters

APPENDIX L: Calculation of Adjustable Parameters


tanks in series) and PeL(Peclet number)
L.1

N (number of

Cakulation sample

The number of tanks in series parameter (N) for the Tank in Series (TIS) model, and the Peclet
number (Pe3 for the Dispersed Plug Flow (PDF)model were calculated for each run of the
tracer tests conducted in both biofilter. The following equations were used.

1. Based on the first moment

where t,, is the mean residence time estimated as the first momentum about the origin of the
retention time distribution function of the tracer cruve, and E(tJ is the exit-age distribution
fhction evaluated at t,,,.
2. Based in the variance (
a
:
)

N=-t

(Eq. L.2)

The calculation of PeL depends on the approach considered when solving the continuity
equation for the DPF model.
Gaussian solution
2t;

Pe, = -

(Eq. L.3)

0:

Solution for an openspen system


6m

Solution for a closed-closed system


2
-(~e,

:.P

: 0
- 1 + e+~)- aL,
=

( ~ q .LS)

Solution for a closed-open system

Effects of Biomuss Growth on Pressure Drop in Biofiters

Appendix L

L-2

L.2 Results
Panmeter calculation based on average tracer c u m 8 from 3 runs
N
Panmeter
P ~ L
hrst moment
Variance
Gauss
Open-open Closed-closed Closed-open
APPmch
11.3
10.9
21.8
THA
25.3
20.8
23.2
March 25,19981 Day 126
22.7
21.6
Nova ln& Biofilter
20.4
21.2
Wood chip biofilter
April15,19981 ay 147
Nova lnert Biofilter
153.0
Wood chip biofitter
April16,1998/ Dey 148
Nova Inert Biofilter
17.9
Wood chip biofilter
18.8

Table L.1

Parameters N and PeLcalculated fiom the tracer curves

Effects of Biomass Growth on Pressure Drop in Biojilrers

Appendix A4

M-1

APPENDIX M: Modelling Pressure Drop as a Function of


Biomass Accumulation Calculations

MODELLINO PRESSURE DROP IN THE NOVA INERT BIOFILTER AS A FUNCTION OF BlOMASS ACCUMULATION
Erimrtad drtm
0.85 '
Sh8pe factor (iphericity)
Avongo prrllck ridius(m)
0.0035 '
SpecHic surface rnr
484
Coordinrtion number
6
Torluonity hctor (rpherer)
1.58 "*
tnitiil permsability(m2)
1.6637t E-07 ***Air vircoaity for 100%RH (kglmr)
9.61 E-06
Air density for 100%RH (kg/m5)
1.O89
'(Sae Appondix J)
**mz prftkle rurlrcel m3 clarn bad substratum
***Ertim8tod with initial porority using Dullien'r axpresaion (Eq. 31)
""Estimited with initial ~ o r o s-tk ~ex~teision
as in Ea. 32
.

Exporlmentrl d r u
Biomarr donsity (kglm')
Initial porority
perimentrl
Cleen bed denrity (kglm5)
(Bulk bed subrtmtum denrity)
Air temperitum (C)

Crois sectionsl are8 (m')


Ergun's constant A
Ergun's constant B
Erpun'r exponent

Section
of the
Bloflltor

D ~ Y

A l i flow

Experlmontal b l o m r r i
concontmUon

BiOIn888
par 8 u d 8 8~r.a
X

monlm
thic knoss
Lf

(Llm in )

1
56 1
2
5

:alculrtia
1.1

1.2

2.1

2.2

3.1

3.2

84
111
148
56
84
111
148
56
84
111
148
56
84
111
148
i urlng

56
84
Ill
148
56
84
111
148
56
84
111
148
56
84
11 1
148
56
84
111
148
56
84
111
148

4.1

56

4.2

84
111
148
56
84
11 1
148

128
130
124
125
128
130
124
125
128
130
124
125
128
130
124

1.853
1.765
2.305
0.838
0.751
0.884
1.447
0.573
0.616
0.61 3
1.250
0.545
0.447
0.678
1.O64

af bl0mi88 conc

itrrtlon d o n g tho bodr


2.582
3.240
2.728
2.888
0.750
1.159
t .283
2.01 4
0.094
0.783
1.281
1.453
0.761
0.735
0.685
1 A45
0.480
0.738
0.600
1.650
0.620
0.554
0.819
1.045
0.447
0.S60
0.844
1.375
0.594
0.381
0.545
0.908

able h( n
based m biomass concentrations in the
Nova Inert biofilter. The resdts that most suitably describe the relation between pressure &op
and biomass concentration in the biofilter are shown in shaded columns.

Effects of Biomass Growth on Pressure Drop in Biofiters

Appendix M

1 . Biofilm growth considering

attached spheres
(Eq. 4 Alonso et al., 1997)

2. Biohlm growtri on
spheres in a cubic arrangement
(Eq. 5 1 Cunningham et al., 1991 )

3. bionlm growth on spheres

with a heterogeneous site


distribution 8 cubic arrangement
(Eq. 12 1 Taylor et al ., 1990)
0.2r10
0.2710
0.27 10
0.27 1O
0.27 10
0.2710
0.27 1O
0.27 10
0.2710
0.2710
0.27 1O
0.2710
0.271O
0.27 10
0.2710
0.27 10

-0.8209
-1.3372
-0.9294
-1 .O503
0.2008
0.0225
-0.0372
-0.4423
0.0992
0.1907
-0.0362
-0.1239
0.1 999
0.21 O8
0.2309
-0.1 193
0.31 13
0.2093
0.2651
-0.2348
0.2572
0.2828
0.2574
0.0760
0.3234
0.2805
0.1213
-0.0833
0.2673
0.3440
0.2863
0.1371

:ompredicting pressru
the Nova Inert biofilter. The results that most suitably describe the relation between pressure
&op and biomass concentration in the biofilter are shown in shaded columns.

Effects of Biomass Growth on Pressure Drop in ~iofiters

Appendix M

M-3

Calcubtion of prerrure drop u8ing Darcy'r Law for gas flow [(APIL)=U~I~
Eq. T)
~ e n e r ~ i i i (KI
ty
I
~ermor~iiity
(11
I
~rsarurecirop (APIL) ~ a a e aon
difterent value8 of permeability
(Eq. 32 1 Au ria, 1993)
(Eq. 21 1 Clement ef al., 1996)

I
l
B. 1

8.2

A. 1

A. 2

B.1

B.2

In2

m2

Palm

Palm

Pdm

Paim

8.18E-09
8.00E-12
8.76E-10
6.94E-09
8.06E-09
9.53E-09
1.27E-08
3.28E-08
9.32E-09
1.Q7E-08

1.79E-08
2.42E-08
1.79E-08
3.77510
3.70E-O8
2.36E-08
f .66E-09
2.02E-08
4.50E-08
2.Sf -O8
2.44E-09

-7.1
-2.8
-5.7
4.3
93.9
101977.3
-25788.3
-28.5
989.6
715.1
-28003.6
-786.9
95.4
81.1
59.5
-873.9
18.7
83.1
35.9
-139.6
38.6
27.7
40.0
2307.4
16.1
28.6
535.5
-2404.3
33.5
12.9
26.9
342.1

:tingpressu
Ion biornass concentrations
the Nova Inert biofilter. The resuk that m% suitably descnbe the relation between pressure
drop and biomass concentration in the biofilter are shown in shaded columns

Effects of Biomass Growth on Pressure Drop in Bioflters

Caltulation of preuure drop urlng the


Modlfied Ergun equalon and difbrsnt value8 of
bed porodty r m crlculrted by modal 1 and 2
1

Paim

txperirnentar
pmsaum drop (APIL)

~aim

from predicting

--

pressure drop based


biomass concentrations in the Nova Inert biofilter. The resuIts that most
suitably describe the relation between pressure &op and biomass concentration
in the biofilter are show in shaded columns

EfJects of Biomass Growth on Pressure Drop t Bioflters

Вам также может понравиться