You are on page 1of 313

CHARACTERISATION AND POTENTIAL USES

OF MICROBIALLY ENHANCED COMPOST


EXTRACTS FOR PLANT GROWTH

PhD THESIS

Submitted in Partial Fulfillment of the


Requirements for the Degree of
Doctor of Philosophy

CENTRE FOR PLANT AND WATER SCIENCE

KARUNA SHRESTHA

MARCH, 2011

Faculty of Sciences, Engineering and Health


CQUniversity, Rockhampton
Australia
ABSTRACT
Microbially enhanced compost extracts are increasingly being used to
improve soil and crop ‗health‘. In this thesis, studies were made both on the process
of preparing compost extracts and on their possible roles in agriculture. Compost
extraction and microbial enhancement processes were characterised for a system
based on open air incubation using chemical, biochemical and molecular techniques,
with variation in additives, dilution, aeration, compost age, compost source including
vermicasts (earthworm casts based on rumen compost) and commercially available
inocula (Living SoilTM and Nutri-Life 4/20TM). The nutritional and microbial
properties of the compost extracts could be improved or manipulated as desired by
supplying additives or aeration in the extract during incubation. Improvement in
community level functional diversity was noted in vermicast leached extract
compared to the original rumen compost, vermicasts and rumen compost extract. The
Biolog and DGGE tools were useful in the overall characterisation of the microbial
assemblages of these liquid samples. The commercial products were comparable in
terms of microbial diversity to compost extracts based on fresh rumen content
compost.
The following hypotheses for the action of compost extracts in improving
crop ‗health‘ and, therefore, their roles in agriculture were tested: (a) direct
improvement of plant nutrition via the nutrient content of the compost extracts; (b)
improvement of plant nutrition through an indirect action, via microbial action in
mineralising of organic or solubilising of inorganic components of the soil; (c)
acceleration of litter decomposition; (d) bio-protection against fungal pathogens
(namely, Fusarium oxysporum, Fusarium solani, and Rhizoctonia solani); and (e)
protection to root growth against high levels of aluminium (Al) ions, via chelation of
Al ions by humic acids or uptake by microbes. The noted increase in tomato and
sorghum plant growth associated with a compost extract treatment (from composted
rumen waste) was attributed to nutrient addition inherent in the treatment rather than
solubilisation of minerals and/or mineralisation of organic matter. Litter
decomposition rates (sugarcane trash mat), as indicated by cumulative respiration
rate, were increased when microbially enhanced compost extract produced from a
mixture of composted rumen waste and sugarcane stubble, was applied. However,
there was no difference between extracts made from composts of different ages. In
vitro studies showed a distinct effect of compost extracts in suppressing fungal wilt
diseases caused by three fungal pathogens (namely, Fusarium oxysporum, F. solani,
Rhizoctonia solani), however, the disease suppressive effect of compost extracts was
not consistent for the three tested pathogens in hydroponic tomato trials. Aluminium
toxicity on maize seedling root growth was alleviated by compost extract application.
A field study indicated that compost extracts enhanced the soil bacterial community
diversity compared to the conventional treatment. It is likely that compost extracts
will find an expanded role in broadscale agriculture and a role as a liquid organic
fertiliser in intensive horticulture.

i
ACKNOWLEDGEMENTS

First and foremost, I would like to thank my principal supervisor, Professor


David J. Midmore, for his keen interest, valuable comments and guidance throughout
my PhD studies.
I wish to express my sincere gratitude and appreciation to Professor Kerry
B. Walsh for his critical comments, valuable suggestions, and guidance. I am very
thankful to my associate supervisor, Associate Professor Keith M. Harrower, who
not only encouraged me to go into microbiology but also taught me the skills
personally in this totally new field to me. Likewise, my sincere thanks go to my
associate supervisor Associate Professor Nanjappa Ashwath for his helpful
discussions and suggestions.
My special thanks go to Rob Lang from Broadmeadows Pty Ltd for
continuously supplying the rumen compost, David Wyatt from Vermicrobe
International Pty LtdTM for donation of vermiculture facilities, William (Brock)
McDonald for the maintenance of the Vermicrobe system and also Professor Andrew
S. Ball and Dr. Eric M. Adetutu for providing the technical and laboratory support at
the Faculty of Science and Engineering of Flinders University, Adelaide. I would
like to acknowledge Ray O‘Grady of O‘Grady Rural (NSW) and Graeme Sait from
NutriTech Solutions (NTS) Pty Ltd (Qld) for supplying their valuable products for
research use. I am also grateful to BSES Ltd, Bundaberg, Qld for supplying the
sugarcane trash material, John Eigens of Pacific Seeds Company for providing maize
seeds for my research and Tom Nicholas for allowing me to accompany him during a
field survey in central and southern Qld and northern NSW to review the current
practices and production of compost extracts in 2008. Moreover, I am highly obliged
to Scott Stevens for providing soil samples from the trial site located in Baralaba on
the central Queensland highlands and treated with different field management
practices including the application of compost extracts.
I express my gratefulness to various reviewers of my thesis for their useful
comments, constructive criticism and invaluable suggestions. I would also like to
thank the anonymous reviewers and editors for their valuable comments on the
revised version of Chapter 2, 3, 4 which have been published in ―Bioresource
Technology‖ and Chapter 7 which has been accepted in ―Biological Control‖.

ii
I would like to extend my sincere gratitude to the Endeavour Postgraduate
Program (EPA) for providing me a scholarship to carry out my PhD studies at
CQUniversity, Australia and I am also grateful to various Endeavour case managers
for resolving the official as well as personal matters throughout my study period. I
am indebted to CQUniversity for their financial support to undertake research studies
during my PhD course.
My sincere acknowledgement goes to Centre for Plant and Water Science
(CPWS) staffs, members, colleagues and friends for their cooperation and support. I
would personally like to thank Dr. Surya P. Bhattarai and Dr. Phul P. Subedi for their
invaluable advice and encouragement. Dr. Alan Keen and Associate Professor Steve
Mckillup from CQUniversity and Christina Playford from DPI are greatly
acknowledged for providing statistical support during the early phase of my studies.
Linda Ahern, indeed, deserves appreciation for her supportive attitude and readiness
to help in any circumstances. Many thanks to Rob Lowry, Jeffrey Conaghan and
Graham Fox for providing technical assistance while undertaking my research.
My husband deserves a special thanks for his endless support, guidance and
assistance in each and every step of my career. I appreciate his patience and deep
understanding throughout my study period. Most importantly, I am thankful to him
for his supervisory role in order to complete my studies.
Last but not least, I would like to extend an earnest thanks to my family,
especially to my parents for their inspiration, love and great support to achieve my
goals; without their encouragement and blessings my journey wouldn‘t have been
successful.

iii
DECLARATION

I hereby declare that this thesis is my original research work and it has not

been submitted anywhere for any award. Contributions of others, if involved, are

either referenced or acknowledged.

I authorise CQUniversity to reproduce this thesis in whole or in part for

purposes of research.

……………………………

Karuna Shrestha

iv
Table of Contents

ABSTRACT ................................................................................................................................i

ACKNOWLEDGEMENTS ........................................................................................................... ii

DECLARATION.........................................................................................................................iv

Chapter 1: A literature review on the use and efficacy of compost extracts in crop production ... 1
1. Overview. ........................................................................................................... 1
2. Soil ‘Health’ ......................................................................................................... 3
3. Organic inputs into agricultural systems ................................................................... 6
4. Composting ......................................................................................................... 8
5. Compost extract ................................................................................................. 12
5.1. Definition of compost extracts ............................................................................................. 12
5.2. Production methods of compost extract .............................................................................. 13
5.3. Defining compost extract quality .......................................................................................... 20
6. Characterising microbial diversity ................................................................. ……..21
7. Current use of compost extract in central and southern Qld and northern NSW .............. 29
8. Compost extract application and claimed benefits ..................................................... 33
8.1. Direct nutrient addition ........................................................................................................ 36
8.2. Soil inorganic fraction solubilisation ..................................................................................... 37
8.3. Soil organic matter decomposition ....................................................................................... 38
8.4. Leaf litter and stubble decay ................................................................................................. 39
8.5. Free living N2 fixation ............................................................................................................ 40
8.6. Plant disease suppression ..................................................................................................... 41
8.7. Acquired systemic resistance ................................................................................................ 47
8.8. Bioremediation...................................................................................................................... 48
9. Summary and hypotheses ..................................................................................... 50
10. Conclusion ......................................................................................................... 53

Chapter 2: Microbial enhancement of compost extracts based on cattle rumen content compost
– characterisation of a system ................................................................................................ 54
Abstract ..................................................................................................................... 54
1. Introduction ....................................................................................................... 55
2. Materials and methods ........................................................................................ 58
2.1. Experiments and treatments ................................................................................................ 58
2.2. Source compost .................................................................................................................... 59
2.3. Physico-chemical analyses .................................................................................................... 60
2.4. Microbial analyses ................................................................................................................ 61
2.5. Statistical analysis ................................................................................................................. 61
3. Results and discussion .......................................................................................... 62
3.1. General trends ...................................................................................................................... 62
3.2. Experiment 1: Comparison of compost extracts with different additives (control, kelp,
molasses and kelp+molasses) ............................................................................................... 63
3.3. Experiment 2: Comparison of compost extracts with different dilution levels .................... 68
3.4. Experiment 3: Comparison of aerated and non-aerated compost extracts with and without
bags ....................................................................................................................................... 72
4. Conclusion ......................................................................................................... 76

Chapter 3: Changes in microbial and nutrient composition associated with rumen


content compost incubation .............................................................................................. 78
Abstract ..................................................................................................................... 78
1. Introduction ....................................................................................................... 79

v
2. Materials and methods ........................................................................................ 81
2.1. Rumen compost .................................................................................................................... 81
2.2. Vermicomposting process, vermicasts and vermicast leachate ........................................... 81
2.3. Physico-chemical analyses .................................................................................................... 82
2.4. Preparation of compost extracts and sampling regimes ...................................................... 82
2.5. Microbial determinations ..................................................................................................... 83
2.6. Microbial biomass carbon analysis ....................................................................................... 84
2.7. Community level physiological profiles of bacteria and fungi .............................................. 84
2.8. PCR and DGGE ....................................................................................................................... 85
2.9. Statistical analysis................................................................................................................. 86
3. Results and discussion .......................................................................................... 86
3.1. Physico-chemical analysis ..................................................................................................... 86
3.2. Microbial and biochemical assays......................................................................................... 87
3.3. Characterisation of compost extracts ................................................................................... 87
3.4. Community level physiological profiles of bacteria and fungi .............................................. 91
3.5. PCR-DGGE analysis ................................................................................................................ 95
4. Conclusions ........................................................................................................ 97

Chapter 4: Comparison of microbially enhanced compost extracts produced from


composted rumen content material and from commercially available inocula ........... 98
Abstract ..................................................................................................................... 98
1. Introduction ....................................................................................................... 99
2. Materials and methods ...................................................................................... 101
2.1. Source compost and commercial preparations .................................................................. 101
2.2. Preparation of compost extracts and commercial extracts ................................................ 102
2.3. Physico-chemical analyses .................................................................................................. 103
2.4. Biochemical and microbial analyses ................................................................................... 103
2.5. Community level physiological profiles .............................................................................. 104
2.6. Microscopic analyses .......................................................................................................... 104
2.7. Genetic diversity of microbial communities of compost extracts ...................................... 104
2.8. Statistical analysis ............................................................................................................... 105
3. Results and discussion ........................................................................................ 105
3.1. Culture characterisation ..................................................................................................... 105
3.2. Microbial and biochemical analyses ................................................................................... 109
3.3. Community level physiological profiles of bacteria and fungi ............................................ 111
3.4. Microscopic analyses .......................................................................................................... 117
3.5. PCR and DGGE ..................................................................................................................... 118
4. Conclusions ...................................................................................................... 122

Chapter 5: Microbially enhanced compost extract: does it increase solubilisation of


minerals and mineralisation of organic matter and thus improve plant nutrition?.. 123
Abstract ................................................................................................................... 123
1. Introduction ..................................................................................................... 124
2. Materials and methods ...................................................................................... 127
2.1. Compost and compost extract ............................................................................................ 127
2.2. Experiment 1 ....................................................................................................................... 128
2.3. Experiment 2 ....................................................................................................................... 129
2.4. Experiment 3 ....................................................................................................................... 131
2.5. Statistical analyses ............................................................................................................. 132
3. Results ............................................................................................................ 133
3.1. Characterisation of compost extract .................................................................................. 133
3.2. Experiment 1: Tomato growth in sand-compost mix ......................................................... 134
3.2.1. Plant biomass .................................................................................................................. 134
3.2.2. Tomato leaf chlorophyll concentration ........................................................................... 135

vi
15
3.2.3. δ N of fertilisers (compost and compost extracts) and plants grown in compost-sand
mix ................................................................................................................................... 136
3.3. Experiment 2: Tomato growth on soil ................................................................................ 137
3.3.1. Soil ................................................................................................................................... 137
3.3.2. Tomato growth, yield and chlorophyll concentration ..................................................... 138
15
3.3.3. δ N of fertilisers and plants grown on soil ..................................................................... 139
3.3.4. Microbial enumeration of aerated and non-aerated compost extracts ......................... 141
3.3.5. Soil microbial populations ............................................................................................... 141
3.3.6. Soil respiration and temperature .................................................................................... 142
3.3.7. Mass balance of nutrients ............................................................................................... 142
3.4. Experiment 3: Sorghum crop rotation ................................................................................ 143
3.4.1. Plant growth, shoot biomass, chlorophyll concentration and light interception ............ 143
4. Discussion ........................................................................................................ 145
4.1. Effects of treatments on growth and nutrition of tomato plants ....................................... 145
4.2. Effects of treatments on tomato leaf chlorophyll concentration ....................................... 147
4.3. Influence of soil types on crop growth ............................................................................... 148
4.4. Nutritional and microbial supply from compost extracts ................................................... 149
4.5. Effects of treatments on soil microbial properties............................................................. 150
5. Conclusions ...................................................................................................... 151

Chapter 6: Biodegradation of sugarcane trash through the use of enhanced compost


extracts.............................................................................................................................. 152
Abstract ................................................................................................................... 152
1. Introduction ..................................................................................................... 153
2. Materials and methods ...................................................................................... 156
2.1. Soil, plant and compost material ........................................................................................ 156
2.2. Biodegradation of trash mat ............................................................................................... 156
2.3. Analytical methods ............................................................................................................. 158
2.3.1. Physico-chemical analyses .............................................................................................. 158
2.3.2. Microbial analyses ........................................................................................................... 158
2.4. Statistical analyses .............................................................................................................. 159
3. Results and discussion ........................................................................................ 159
3.1. Physico-chemical, biochemical and microbial analyses of different aged composts ......... 159
3.2. Characterising the incubation process using different aged compost extracts as inoculum161
3.3. Comparison of chemical, microbial and biochemical characteristics of different aged and
sugarcane stubble amended compost (24 h incubated) extracts ....................................... 164
3.4. Effects of compost extract treatments on sugarcane trash degradation ........................... 167
3.5. Weight loss of trash and cumulative respiration rate ......................................................... 170
4. Conclusions ...................................................................................................... 174

Chapter 7: Suppressing fungal wilt diseases of tomato caused by Fusarium


oxysporum, F. solani and Rhizoctonia solani using compost extracts in vitro and in
hydroponics culture ......................................................................................................... 175
Abstract ................................................................................................................... 175
1. Introduction ..................................................................................................... 176
2. Materials and methods ...................................................................................... 179
2.1. In vitro hyphal growth trials ................................................................................................ 180
2.2. Hydroponics trials ............................................................................................................... 182
2.3. Statistical analyses .............................................................................................................. 184
3. Results and discussion ........................................................................................ 185
3.1. In vitro trials ........................................................................................................................ 185
3.1.1. Mycelial growth of fungi .................................................................................................. 185
3.1.2. Spore count ..................................................................................................................... 188
3.1.3. Total fungal biomass........................................................................................................ 190
3.2. Hydroponics trials ............................................................................................................... 192

vii
4. Conclusions ...................................................................................................... 199

Chapter 8: The effect of compost extract on root elongation of maize (Zea mays L.)
in the presence of aluminium .......................................................................................... 200
Abstract ................................................................................................................... 200
1. Introduction ..................................................................................................... 201
2. Materials and methods ...................................................................................... 205
2.1. Rumen compost extract and vermicast leachate ............................................................... 205
2.2. Chemical and microbial analyses of compost extracts ....................................................... 205
2.3. Plant material and experimental set up ............................................................................. 206
2.4. Statistical analysis ............................................................................................................... 207
3. Results and discussion ........................................................................................ 208
3.1. Effect of Al on root growth of maize seedlings - Preliminary trial ...................................... 208
3.2. Composition of compost extract ........................................................................................ 209
3.3. Experiment 1 ....................................................................................................................... 210
3.4. Experiment 2 ....................................................................................................................... 212
4. Conclusion ....................................................................................................... 214

Chapter 9: Enhancing microbial activity of a field soil with long-term use of liquid
biofertilisers...................................................................................................................... 215
Abstract ......................................................................................................................................... 215
1. Introduction ........................................................................................................................ 216
2. Materials and methods ....................................................................................................... 219
2.1. Study site and treatments................................................................................................... 219
2.2. Soil sampling ....................................................................................................................... 222
2.3. Soil characterisation ........................................................................................................... 223
2.4. Statistical analyses .............................................................................................................. 224
3. Results ................................................................................................................................. 225
3.1. Physico-chemical, microbial and biochemical analyses ...................................................... 225
3.2. Community level physiological profiles of bacteria and fungi ............................................ 229
3.3. PCR-DGGE analysis for bacteria and fungi .......................................................................... 233
4. Discussion ........................................................................................................................... 237
4.1. Seasonal variation ............................................................................................................... 237
4.2. Comparison of brigalow and cultivated soils ...................................................................... 238
4.3. Effects of management practices on soil properties .......................................................... 240
5. Conclusions ......................................................................................................................... 241

Chapter 10 ........................................................................................................................ 243


Summary and future directions ..................................................................................... 243
References ........................................................................................................................ 250

viii
List of Tables
Chapter 1
TM
Table 1. 1. Carbon substrates of Ecoplate and FF microplate (Biolog ), respectively divided
into different substrate guilds according to Preston-Mafham et al. (2002) ................................ 25
Table 1. 2. List of farmers [permissions granted to use their data] from Qld and NSW .............. 30
Table 1. 3. List of companies [permissions granted to use their data] visited with Tom Nicholas
or contacted that were involved in producing compost extract .................................................. 32
Table 1. 4. Summary of reports on proven disease suppression following application of
compost extracts (CEs) produced from known input materials................................................... 44
Chapter 2
Table 2. 1. Comparison of physico-chemical and microbial characteristics of compost extracts
incubated with different additives ............................................................................................... 67
Table 2. 2. Comparison of physico-chemical and microbial characteristics of compost extracts
incubated with different dilutions ............................................................................................... 69
Table 2. 3. Comparison of physico-chemical and microbial characteristics of compost extracts
incubated for 24 h in presence and absence of aeration ............................................................. 75
Chapter 3
Table 3. 1. Initial physico-chemical, microbial and biochemical characteristics of rumen
compost and vermicast (n = 3). .................................................................................................... 87
Table 3. 2. Comparison of chemical, biochemical and microbial characteristics of rumen
compost extract (RCE), vermicast leached extract (VLE), vermicast extract (VCE) ...................... 91
Table 3. 3. Shannon weaver Diversity Index (H’), Equitability Index (J) and number of bands (S)
for bacteria and fungi present in rumen compost (RC), vermicast (VC), rumen compost extract
(RCE) and vermicast leached extract (VLE) .................................................................................. 96
Chapter 4
Table 4.1. Comparison of physico-chemical and microbial characteristics of different liquid
cultures sampled at 24 h and at 48 h of incubation. .................................................................. 106
Table 4.2. Sample analysis data of nine month old rumen compost extract (9MCE) and LS
TM
(Living Soil ) .............................................................................................................................. 117
Table 4.3. Diversity and equitability of microbial communities in different liquid cultures...... 120
Chapter 5
Table 5. 1. Nutrient (macro and micro) concentrations for aerated and non-aerated compost
extract samples analysed by Wesfarmers CSBP Limited, Australia............................................ 133
Table 5. 2. Characterisation of vertisol and ferrosol used in the greenhouse trial ................... 138
Table 5. 3. Enumeration of bacteria and fungi populations in representative samples of ACE
and NCE. The compost extracts were produced at the ratio of 1:10 w/v .................................. 141
Table 5. 4. Effect of compost extracts on microbial (bacterial and fungal) populations (cfu Log
-1
10 mL ) in two soil types sampled at 36 days after transplanting ............................................. 142
Table 5. 5. Mass balance of N in compost extract and tomato crop grown in vertisol. ........... 143
Chapter 6
Table 6. 1. Physico-chemical, biochemical and microbial comparisons of different aged, and
sugarcane stubble amended, composts .................................................................................... 161
Table 6. 2. Comparison of chemical, biochemical and microbial characteristics of different
aged and sugarcane stubble amended compost extracts incubated for 24 h with 1% kelp and
0.5% molasses as amendments ................................................................................................. 165
Table 6. 3. Chemical, biochemical and microbial characteristics of soil samples (collected after
168 days of incubation). ............................................................................................................. 169

ix
Chapter 7
Table 7. 1. Fusarium infection in roots based on rating scale of 0 (pathogen absent) to 1
(pathogen present) on eight tomato plant roots in each hydroponics container ..................... 194
Chapter 8
Table 8. 1. Comparison of physico-chemical and microbial characteristics of sterilised and non-
sterilised rumen compost extract and vermicast leachate ........................................................ 209
Chapter 9
Table 9. 1. Details of treatments and fertility programs of a long-term field trial (commenced
in 2007) conducted in Baralaba, Qld .......................................................................................... 221
Table 9. 2. Variation in the physico-chemical, microbial and biochemical characteristics of soil
samples collected from plots treated with different management practices at the Baralaba trial
site in two different seasons ...................................................................................................... 228
Table 9. 3. Shannon-Weaver Diversity Index (H’), Equitability Index (J) and number of bands (S)
for bacteria and fungi present in soil samples collected in the dry-winter season from plots
treated with different management practices ........................................................................... 236

x
List of Figures

Chapter 1

Figure 1. 1. Enzymatic conversion of FDA to fluorescein (Source: Adam & Duncan, 2001) ........ 22
Figure 1. 2. Flow diagram showing the different steps in the analysis of microbial community
structure by PCR-DGGE ................................................................................................................ 28

Chapter 2
-1
Figure 2. 1. Comparison of pH, electrical conductivity (dS m ), fluctuations in temperatures
(˚C) and dissolved oxygen (% of saturation) over time (readings taken at every 3 h intervals
until 24 h and final readings taken at 48 h) in rumen compost extracts with different additives65
Figure 2. 2. Total microbial activity over time (readings taken at 12, 24 and 48 h) as estimated
by the FDA hydrolysis method in compost extracts with different additives .............................. 67
-1
Figure 2. 3. Time course of solution pH, electrical conductivity (dS m ), temperatures (˚C) and
dissolved oxygen (% of saturation); readings taken at every 3 h intervals until 24 h and final
readings taken at 48 h in rumen compost extracts with different dilutions. ............................... 70
Figure 2. 4. Comparison of total microbial activity at different times (readings taken at 12, 24
and 48 h) following the FDA hydrolysis method in compost extracts with different dilutions .... 71
-1
Figure 2. 5. Comparison of pH, electrical conductivity (dS m ), fluctuations in temperatures
(˚C) and dissolved oxygen (% of saturation) over time ................................................................ 73
Figure 2. 6. Comparison of total microbial activity over time (readings taken at 12, 24 and 48 h)
following the FDA hydrolysis method in compost extracts in presence and absence of aeration
with and without the use of bags. ............................................................................................... 75

Chapter 3
-1
Figure 3. 1. Comparison of pH, EC (dS m ), fluctuations in temperatures (°C) and dissolved
oxygen (%) per unit time (readings taken at every 3 h intervals until 24 h after commencement
of compost extract incubation ..................................................................................................... 89
Figure 3. 2. Comparison of total microbial activity per unit time (readings taken at 12, 24 and
48 h) following the FDA hydrolysis method ................................................................................. 91
Figure 3. 3. Kinetics of average well colour development (AWCD) and optical densities of (A)
bacterial (570 nm) and (B) fungal (490 nm) communities. .......................................................... 93
Figure 3. 4. Principal component analysis of optical densities produced by microbial activities
of rumen compost (RC), vermicast (VC), rumen compost extract (RCE) and vermicast leached
extract (VLE) ................................................................................................................................. 94
Figure 3. 5. UPGMA (Unweighted Pair Group Method with Arithmetic mean algorithm)
dendrogram generated from responses to 16S rDNA and Internal Transcribed Spacer region
based Denaturing Gradient Gel Electrophoresis (DGGE) analyses .............................................. 96

Chapter 4
-1
Figure 4.1. Comparison of (A) pH, (B) electrical conductivity (dS m ), (C) temperatures (˚C) and
(D) dissolved oxygen (ppm) over time (readings taken at every 3 h intervals until 24 h and final
readings taken at 48 h) in different cultures .............................................................................. 107
Figure 4.2. Comparison of total microbial activity at different times (readings taken at 12, 24
and 48 h) following the FDA hydrolysis method in different cultures ....................................... 110
Figure 4.3. Kinetics of average well colour development (AWCD) and optical densities of (A)
bacterial (570 nm) and (B) fungal (490 nm) communities, respectively, from different compost
extracts and microbial preparations. ......................................................................................... 112
Figure 4.4. Average carbon substrate utilisation from different substrates by samples from
three month old rumen compost extract (3MCE), nine month old rumen compost extract
TM
(9MCE), nine month old rumen compost extract plus Nutri-Life 4/20 inoculum (9MTCE),
TM TM
Living soil (LS) and Nutri-Life 4/20 (NL-4/20) ...................................................................... 114

xi
Figure 4.5. Principal component analysis (PCA) of optical densities produced by microbial
activities of five different cultures consuming 32 (Biolog Ecoplate) and 96 (Biolog FF
microplate) different food sources including the control .......................................................... 116
Figure 4.6. UPGMA (Unweighted Pair Group Method with Arithmetic mean algorithm)
dendrogram for (A) bacteria and (B) fungi ................................................................................. 119

Chapter 5

Figure 5. 1. Effect of compost extracts on (A) total plant biomass (g d wt) and (B) shoot: root
biomass of tomato plants in Experiment 1 grown in a thoroughly washed compost-sand mix at
1: 5 ratio in two replicate trials .................................................................................................. 135
Figure 5. 2. Effect of compost extracts in Experiment 1 on chlorophyll concentration in leaves
in (A) trial I, and (B) trial II .......................................................................................................... 136
15
Figure 5. 3. δ N values of fertilisers applied including the rumen compost and of the tomato
plants grown in compost-sand mix (1: 5) under growth chamber conditions ........................... 137
Figure 5. 4. Data from Experiment 2 on (A) total above-ground biomass (g d wt per plant) of
tomato subject to four treatments after four months of planting on two soil types, (B) fruit
yield of tomato (g d wt per plant), (C) leaf chlorophyll concentration (SPAD unit) ................... 139
15
Figure 5. 5. δ N values of tomato plants grown in ferrosol conducted in a greenhouse
(Experiment 2)............................................................................................................................ 140
Figure 5. 6. Data on sorghum from Experiment 3 on (A) above-ground dry weight biomass per
plant, (B) chlorophyll concentrations of leaves and (C) percent light interception ................... 144

Chapter 6
-1
Figure 6. 1. Comparison of pH, electrical conductivity (dS m ), fluctuations in temperatures
(˚C) and dissolved oxygen (% of saturation) over time (readings taken at every 3 h intervals
until 24 h and final readings taken at 48 h) ............................................................................... 163
Figure 6. 2. Comparison of total microbial activity at different times (readings taken at 12, 24
and 48 h) following the FDA hydrolysis method in different aged (one, three, six and ten
month old) and sugarcane stubble amended compost extracts ............................................... 166
Figure 6. 3. On the left (A) showing sugarcane trash treated with water control and on the
right (B) showing the same trash treated with compost extract ............................................... 171
Figure 6. 4. (A) Time course of respiration rates which was assessed as accumulation of CO 2
over a 24 h period and (B) comparison of cumulative respiration rates of sugarcane trash
materials lying on top of the ferrosol soil over six month (28 weeks) ....................................... 173

Chapter 7
-1
Figure 7. 1. Effect of compost extracts in trial 1 on mycelial growth rates (mm d ) of Fusarium
oxysporum, f. sp. lycopersici six days after inoculation.............................................................. 186
-1
Figure 7. 2. Effect of compost extracts in trial 2 on mycelial growth rates (mm d ) of Fusarium
oxysporum, f. sp. lycopersici, F. solani DAR 66102, and Rhizoctonia solani DAR 61830 over (A)
three and (B) six days after inoculation .................................................................................... 187
Figure 7. 3. Effect of compost extracts in trial 1 on sporulation (number of spores per plate) of
Fusarium oxysporum, f. sp. lycopersici six days after inoculation .............................................. 188
Figure 7. 4. Effect of compost extracts in trial 2 on sporulation (number of spores per plate) of
Fusarium oxysporum, f. sp. lycopersici and F. solani DAR 66102 ............................................... 189
Figure 7. 5. Effect of compost extracts in trial 1 on total biomass (d wt) of Fusarium oxysporum,
f. sp. lycopersici six days after inoculation ................................................................................. 190
Figure 7. 6. Effect of compost extracts in trial 2 on total biomass (d wt) of Fusarium oxysporum,
f. sp. lycopersici (black bar), F. solani DAR 66102 (light grey bar), and Rhizoctonia solani DAR
61830 (dark grey bar), six days after inoculation ....................................................................... 192
Figure 7. 7. (A) A graph showing an effect of application of compost extracts in hydroponics
trial 3 on two month old tomato plants against the severity of fungal wilt caused by F.
oxysporum f. sp. lycopersici grown in the greenhouse (B) pictures of pathogen inoculated
control plants ............................................................................................................................. 193

xii
Figure 7. 8. Effect of application of compost extracts in hydroponics trial 4 on two month old
tomato plants against the severity of fungal wilt diseases caused by (A) F. oxysporum, (B) F.
solani and (C) R. solani in greenhouse assays ............................................................................ 196
Figure 7. 9. Effect of application of compost extracts in hydroponics trial 4 on the leaf
chlorophyll concentration of two month old tomato plants grown in the greenhouse and
inoculated with three different fungal pathogens ..................................................................... 198

Chapter 8

Figure 8. 1. Relative root extension (% of control, where control plants exhibited a root
extension of 14 cm) of maize (Zea mays L. cv. Hycorn 675 IT) .................................................. 208
Figure 8. 2. Experiment 1: Root elongation rate of maize (Zea mays L. cv. Hycorn 675 IT)
seedlings (n = 4) following growth for 96 h on a solution culture with or without 200 µM of Al
at pH 4.5, with treatments of non-sterilised and sterilised rumen compost extracts and
vermicast leachate (1:500 dilution) ........................................................................................... 210
Figure 8. 3. Experiment 2: Root elongation rate of maize (Zea mays L. cv. Hycorn 675 IT)
seedlings (n = 4) measured 96 h after imbibition in solution culture in the presence or absence
of 200 µL Al (solution pH 4.5) ..................................................................................................... 213

Chapter 9

Figure 9. 1. Kinetics of average well colour development (AWCD) of (A) bacterial (570 nm) and
(B) fungal (490 nm) communities originating from soil samples collected from the trial site
during the dry-winter season at Baralaba .................................................................................. 230
Figure 9. 2. Average carbon substrate utilisation for different substrate categories in terms of
(A) 32 (Biolog Ecoplate) and (B) 96 (Biolog FF microplate) different food sources ................... 231
Figure 9. 3. Principal components analysis of optical densities produced by microbial activities
of soil samples collected from the trial site at Baralaba ............................................................ 232
Figure 9. 4. Unweighted Pair Group Method with Arithmetic mean algorithm dendrogram
constructed from the DGGE profiles of (A) bacterial 16S rDNA and (B) fungal ITS genes
amplified from duplicate DNA samples of soils ......................................................................... 234

xiii
Chapter 1: Literature review

Chapter 1

A literature review on the use and efficacy of

compost extracts in crop production

1. Overview

The large increase in world agricultural productivity during the 1960s and

70s (the ‗green revolution‘) was supported by improved genetics and by

increased chemical inputs (inorganic fertilisers, herbicides and pesticides).

However, there is a current trend to minimise (or exclude, in the case of ‗organic‘

production systems) chemical inputs, on the basis of reducing harm to the

environment and, in some cases, to reduce input costs. A key element of this

trend is a focus on soil ‗health‘, with attention to biological components (e.g. ‗soil

food web‘) as well as physico-chemical attributes (e.g. soil aeration, cation

exchange capacity) (Elfstrand et al., 2007; Supanjani et al., 2006).

A wide array of formulations and products are now marketed in

developed countries in support of this focus on soil ‗health‘. Typically, these

products are biological amendments that are applied with the aim of restoring and

regenerating soil microbial communities, with typical claims including action as a

bio-fertiliser, a bio-decomposer and/or a disease suppressor (Antonio et al., 2008;

1
Chapter 1: Literature review

Elliott & Lynch, 1994; Girvan et al., 2004; Grayston et al., 2004; Hall et al.,

2006; Liu et al., 2007). However, documentation of the efficacy, and more

specifically the mechanism of action, of these products is typically very poor.

One class of such products are ‗compost extracts‘, which are commonly

known as ‗compost tea‘. Production of ‗compost extract‘ has progressed from a

simple, anaerobic extraction process to an ‗aerobic bio-fermenter‘ approach,

expanding from small, simple on-farm activity (cow manure soaking in a 44

gallon drum) to large, relatively elaborated activity (e.g. specialised incubating

and aeration apparatus, addition of microbial food sources, enhancement with

specific microbial strains, centralised production with distribution to farms, and

sophisticated distribution/application systems on farm). The term ‗microbially

enhanced extract‘ is thus technically more correct, and the term ‗compost extract‘

is thus simplistic. However, for convenience, the term ‗compost extract‘ (CE)

will be used in this thesis.

Compost is bulky and hence is costly to transport. Use in broad-acre

production is, therefore, challenging and often replaced by ‗green manuring‘ (i.e.

growth of a crop to be left in the field to decompose). Compost extract can be

seen as a value-add to compost. As a liquid, it is easily applied. Further, compost

can only be applied to the soil, while compost extract can be applied to the soil,

foliage and plant remains, either by drenching or by spraying (Litterick et al.,

2004; Scheuerell & Mahaffee, 2004; Weltzein, 1992; Yohalem et al., 1994).

Compost extracts are gaining popularity, particularly, among those who

are seeking substitutes to agrochemicals (Bess, 2000; Touart, 2000). Organic

farmers are paying more attention to compost extract because of its claimed

disease suppressive properties and the lack of standard disease management tools

2
Chapter 1: Literature review

in the organic sector (Scheuerell & Mahaffee, 2002; Siddiqui et al., 2009).

‗Mainstream‘ growers are also looking for alternative approaches to the use of

petrochemical-derived fertiliser products, as prices soar due to increasing global

demand and limited supply. As an example, the US Department of Agriculture

reported a 65% increase in price of synthetic fertiliser, such as nitrate, potassium

and phosphate over 12 months from date to date (Etter, 2008).

However, there is relatively little verification of these claims

within the scientific literature (as reviewed later in this thesis). This situation is

likely due, in large part, to the variation in the product (effectively each compost

extract is a unique biological community), to the scope of the claims (from

disease suppression to soil phosphate solubilisation), and to the time frame that

may be required (multiple years).

The aim of this review is to summarise reports on the efficacy and mode

of action of compost extracts in agriculture.

2. Soil ‘Health’

Soil health has been defined as the capability of soil to act as a living

system, sustaining life (plants, animals including microorganisms) and

maintaining air and water quality (Doran & Zeiss, 2000). Ask a farmer to define a

‗healthy soil‘, and she/he will speak of the feel and smell of the soil.

Plants are autotrophic, that is, they require only sunlight, CO2, water and

the essential mineral elements. In hydroponic systems these mineral requirements

are typically completely met by supplying them in a soluble inorganic form. In

soil, there is a reserve pool of these minerals in both inorganic and organic

3
Chapter 1: Literature review

insoluble forms. Soil biota are essential to functions of mineralisation of organic

matter and solubilisation of inorganic minerals.

Plants have evolved functional linkages with certain groups of soil biota.

For example, the mycorrhizal symbiosis that exists between plants and certain

ascomycete and basidiomycete fungi is well established to improve plant access

to soil nutrients, particularly P (Finlay & Rosling, 2006; Nehls, 2008). The

functional linkage between some plant species and certain N2 fixing bacteria is

also well established. This linkage involves a tight symbiosis in some cases, for

example, legumes with rhizobia, casuarina with frankia, cycads and azolla with

blue green algae, and a looser commensal relationship in other cases (e.g.

enhanced populations of free living N2 fixers, such as Azotobacter and

Azospirillum are examples of microbiota found in the rhizosphere or phylloplane)

(Kahindi et al., 1997; Unkovich et al., 2008).

Other soil microbiota are not directly associated with plants, but

nevertheless contribute to processes that favour plant growth. Examples of

organisms that contribute to mineral solubilisation in the soil include the

phosphate solubilising bacteria (Ekin, 2010; Rodríguez & Fraga, 1999).

Examples of organisms that contribute to organic matter breakdown, and thus

nutrient availability through mineralisation include bacteria (e.g. Nitrococcus,

Nitrobacter), fungi (e.g. Rhizopus, Fusarium spp.), flagellates, amoebae and

nematodes (Bloem et al., 1994; Bourne et al., 2008). Other microbiota are

considered beneficial through an antagonistic effect on the growth of pathogenic

organisms (e.g. Dubeikovsky et al., 1993; Hoitink et al., 1997; Siddiqui et al.,

2008a).

4
Chapter 1: Literature review

Thus, soil biota are important to ecosystem function, and to that subset of

ecosystems, agricultural systems. One cause of decreased crop yields over time,

that requires increasingly greater chemical inputs, may be farming practices that

have led to a decrease in soil biota. For example, it is concluded that ‗long fallow

disorder‘, wherein a poor crop yield is obtained after a long fallow period, is

linked to a loss of soil biota (Thompson, 1996; Thompson, 1987). Heavy use of

inorganic fertilisers and synthetic biocides are believed to result in reduced

populations of beneficial soil microorganisms, for example, by creating excess

salt and nitrates in the soil (Abbott & Murphy, 2003; Lampkin, 1990). This loss

of soil microbiota is considered to result in a greater dependency on the chemical

inputs.

However, while loss of microbial diversity, is generally associated with a

decline in soil ‗health‘ (Swift et al., 1996), this is not always so. Further, it is

necessary to separate ‗cause‘ and ‗effect‘ between levels of soil microbiota

(density and diversity) and soil fertility/plant growth.

For example, no difference was reported in soil microbial biomass and

microbial diversity between soils differing in intensity of use (e.g. Jesus et al.,

2009; Wardle et al., 1999). Jesus et al. (2009) analysed soil samples (0-20 cm

depth) collected from sites cultivated with crops (maize, banana, cassava,

pineapple and sugarcane), pasture, agroforesty (peach palm, cupuassu, pineapple,

banana and coffee), young and old secondary forest (aged 5 and 5-30 years old,

respectively) and compared with the primary forest, all of which were highland

forest originally. They found that the bacterial diversity, assessed by terminal

restriction fragment length polymorphism (T-RLFP) followed by cloning and

sequencing, did not reduce when forest was converted to pasture and crops, or

5
Chapter 1: Literature review

when cleared land was returned to forest. In fact, they found more diverse

communities in pasture compared to the primary forest. Wardle et al. (1999)

investigated three levels of agricultural intensification, namely, (i) least intensive

cultivation (i.e. hand-hoeing), (ii) repeated cultivation (i.e. cultivating between

the maize rows for weed control) and (iii) herbicide application (rimsulfuron

followed by primisulfuron) and mulching (with 10 cm thick layer of non-treated

pine sawdust) in two sites. They did not find much variation in microbial biomass

and activity between treatments, although mulching was noted to improve soil C

compared to other treatments in both sites.

A recent interesting development of a formal assessment system for soil

‗quality‘ has been the Grains Research and Development Corporation (GRDC)

supported development of a website (www.soilquality.org.au). It allows

comparison of input data of physical (e.g. bulk density), chemical (e.g. lime

benefit), and biological (e.g. organic matter biomass, yield potential) properties of

a given region. While currently established for Western Australia, the intent is to

extend this service to the cropping lands of other Australian states. The characters

used to define soil biology are total organic C, labile C, microbial biomass C, soil

microbial activity (as indexed by soil respiration), soil N supply, presence of

pathogenic fungi Rhizoctonia or take-all, and nematode populations.

3. Organic inputs into agricultural systems

During the past century, the invention of synthetic fertilisers, herbicides

and pesticides has revolutionised the farming industry. However, indiscriminate

use of ‗modern‘ management practices has had negative consequences on the

ecology of agriculture (Danielle & Rai, 2006).

6
Chapter 1: Literature review

Soil organic amendments within cultivated agricultural ecosystems have a

strong influence on the overall structure and function of soil microbial

communities (Buckley & Schmidt, 2001; Steenwerth et al., 2003), resulting in a

higher mineralisation and solubilisation capacity than for a soil to which mineral

fertiliser only is applied (Gerhardt, 1997; Liang et al., 2005; Marcote et al., 2001;

Melero et al., 2006; Pascual et al., 1998; Rao & Pathak, 1996). However, the

influence of these inputs depend largely upon type and quantity of amendment

(Barzengar et al., 2002; Nelson & Oades, 1998). For example, while the likely

impact of large quantities of green manure or compost (representing both a

microbial food and inoculum source) to a soil on soil microbiota levels and

diversity is obvious, the impact of a small volume of a microbial culture over a

large area of land is less obvious.

Organic farming involves crop production without using synthetic

chemicals or most inorganic fertilisers (Winter & Davis, 2006). Proponents of

organic farming believe that organic farming is environmentally benign and

organic products have increased nutrient content and can improve human health

(Bourn & Prescott, 2002). Organic farming is highly reliant on the biological

fertility of the soil as a determinant of soil physical and chemical fertility (Abbott

& Murphy, 2003; Bulluck et al., 2002; Ferreras et al., 2006). The crux of this

farming practice is thus ―soil health‖, and a major theme of organic agriculture is

to build a diverse microbial community in the expectation of maximising nutrient

use efficiency in the soil (Giller et al., 1997; Janvier et al., 2007; Moeskops et al.,

2010). Management practices, such as addition of organic matter, maintaining

plant diversity, reduced tillage and adding soil amendments may benefit soil

biology and are commonly employed to improve soil health (Berner et al., 2008;

7
Chapter 1: Literature review

Cotxarrera et al., 2002; Erhart et al., 1999; Lee & Pankhurst, 1992; Pe´rez-

Piqueres et al., 2006; Rincon et al., 2006). These practices enhance nutrient

availability, soil physical structure and soil microflora (Canullo et al., 1992;

Midmore, 2006; van Elsas et al., 2002). The use of compost extract is another

potential management tool.

4. Composting

Compost extract quality is dependent on the quality of compost used

(Carpenter, 2005). Composting can be viewed as an organic waste reduction

process (Eiland et al., 2001), but its greater value lies in the use of compost as a

soil conditioner containing soluble mineral nutrients, organic matter, including

humic acids, and beneficial microorganisms (Leu, 2006). Compost quality is

typically determined by the absence of viable weed seeds and disease-causing

organisms (Jones & Wyatt, 2005), by its nutrient content (e.g. soluble and total

N) and by its physical structure. The Australian standard for compost quality lists

the following specifications: pH 5.0-7.5; EC 0-1 dS m-1; organic matter content ≥

25% dry matter; total N ≥ 0.8% dry matter; total P ≤ 0.1% dry matter, if for P

sensitive plants, and soluble P ≤ 5 mg L-1, if for P tolerant plants and C:N ≤ 22

(AS4454, 1999). While microbial diversity and total microbial load is a quality

factor, it is not assessed.

Composting is a process of usually aerobic decomposition of organic

matter by microbiota, producing a stable organic end product, with carbon-

dioxide, water and heat as by-products (Rynk et al., 1992). This process is an

extension of that occurring in natural ecosystems (e.g. forest floor leaf litter

decomposition). The major factors that influence compost quality include

8
Chapter 1: Literature review

materials added (C/N ratio, particle size, fibre and lignin content), and

temperature, aeration, pH, and moisture (Avnimelech et al., 2004). The main

biota in compost include bacteria, fungi, protozoa, nematodes and rotifers

(Ingham, 2003).

As a generalisation, fungi will dominate, relative to bacteria, in compost

containing a relatively high proportion of cellulose and lignin (Pe´rez et al., 2002;

Tuomela et al., 2000). Typical bacterial genera present in compost include

Azotobacter, Nitrobacter, Azospirillum, Pseudomonas, Proteobacteria,

Actinobacteria and so on (Tang et al., 2007). Typical fungal genera present in

compost include Fusarium and Trichoderma.

The windrow or pile system is the most common form of composting. The

recommended size for manure based windrow is 1.5 m in height and 2.5 m wide

at the base (Rynk et al., 1992). Composting typically involves three different

phases, namely, (i) mesophilic, (ii) thermophilic and (iii) maturation, the first

being a moderate temperature phase (0-45˚C) that lasts for a few days, the second

represents a high temperature phase (50-60˚C or above) that persists for a few

days to several months, and the last, a cooling phase (to ambient temperature)

that lasts for several months (Rynk et al., 1992). During the second phase, when

temperatures are typically at or over 55˚C, most weed seed and human and plant

pathogenic microorganisms are destroyed (Grundy et al., 1998; Suarez-Estrella et

al., 2003). For good compost production, the temperature should be maintained at

57-68˚C, and always below 70˚C, during this phase (Rynk et al., 1992). High

temperatures, greater than 70˚C result in a decline in microbial activity (OSUE,

1999).

9
Chapter 1: Literature review

Aeration is typically achieved by structural means (e.g. inclusion of

coarse organic materials into the pile to provide aeration pathways) and by

practices, such as frequent turning of the pile (Tiquia, 1996). Compost

temperature, an index of aerobic decomposition, is monitored to determine the

frequency of aeration (Michel et al., 1996a; Tiquia, 1996). Turning frequency

also depends on decomposition rate, porosity and moisture of the composted

materials (Tiquia et al., 1996), and the size of the pile (Michel et al., 1996b).

During the thermophilic phase, daily turnings may be required for easily

decomposable mixes high in nitrogen, while the frequency of turning decreases,

for example, to weekly or fortnightly turnings as the pile matures (Misra et al.,

2003; Rynk et al., 1992). The windrow pile is usually turned if temperatures fall

below 48˚C so as to stimulate aerobic decomposition (Misra et al., 2003). If

compost dominated by fungal growth is desired, less turning is required than for

bacterial compost as fungal growth is inhibited by frequent turning (Rynk et al.,

1992).

Aerobic composting occurs at a broad range of pH (5.5-9.0), however the

preferred range for pH is 6.5-8.0 (Rynk et al., 1992). The ideal bulk C:N ratio for

active composting is 25:1 to 30:1; with ranges as low as 20:1 acceptable (Rynk et

al., 1992). If this ratio is above 40:1, a longer composting period is required, and

there is the risk that, on application, the compost will immobilise soil N. If the

ratio is too low, nitrogen will be lost in the form of ammonia, resulting in

unpleasant odours.

Anaerobic composting, e.g. Japanese composting system is an efficient

way to decompose organic wastes using microbial inoculants, such as species of

lactic acid bacteria, photosynthetic bacteria, yeasts, and Actinomycetes,

10
Chapter 1: Literature review

producing ―bokashi‖ as an end product (Xu, 2000). This composting makes use

of anaerobic (oxygen free) process that relies on fermented enzymes for enhanced

decomposition of organic materials in one or two weeks. It produces no odours,

enhances crop growth and restores the microbial activity and fertility in heavily

depleted or nutrient poor soils.

Generally, the moisture content in composting materials should be

maintained between 40-65% (Misra et al., 2003). Moisture is necessary not only

for supporting microbial activity and metabolism but also for nutrient transport

(Shi et al., 1999). A moisture level below 45% results in slowed microbial

activity whereas at moisture levels above 65% pore spaces are occupied with

water making the process anaerobic (Rynk et al., 1992). Post production, compost

should not be over dried (i.e. less than 45% moisture) because microbial activity

will be lost, dust will be generated while handling, and it will be difficult to re-

wet the compost (Moon, 1997).

Inadequate protection and improper storage (e.g. open air storage) of

compost can subsequently result in pathogen contamination (Deportes et al.,

1995). Various groups of pathogens may exist in composted raw materials of

various origins (e.g. manure, sewage sludge, green waste and municipal solid

wastes). Such pathogens are mainly of faecal origin (Deportes et al., 1998), and

the group of pathogens such as E. coli, Clostridium botulinum, Salmonella,

Bacillus cereus, and Classical Swine Fever Virus are mainly responsible for

spreading human and animal diseases (Jones & Martin, 2003).

Effectively any organic material can be composted, although ‗home‘

producers generally avoid animal flesh and fats due to the potential to attract

other animals, for fly strike and for development of off-odours. The key to

11
Chapter 1: Literature review

consistent compost quality is consistent inputs. Given compost is a relatively low

value and bulky material, production from locally available materials is essential

to minimise transportation costs. Typical materials are plant-based (leaves, bark,

crop residues, grass clippings, paper wastes and food processing wastes) and

animal based (cattle manure, horse manure, poultry manure, fish processing

wastes, sewage sludge, slaughterhouse wastes).

Raw input material is typically screened to remove unwanted material,

and shredded to reduce particle size. The final compost is also often screened, to

remove clumps of compost, stones and uncomposted materials (Rynk et al.,

1992).

Unlike a chemical fertiliser, compost is not static for it contains live

microbes (bacteria, fungi, nematodes, protozoa). Thus, physical and chemical

properties of compost change continuously. Generally, plant-based composts are

ready in two to six months. Well aged compost will have microbial populations in

a plateau growth phase, and thus in a relatively static phase (Reeve et al., 2010).

Commercial sale of compost (e.g. bagged compost in retail stores) should be well

aged compost.

5. Compost extract

5.1. Definition of compost extracts

‗Compost extract‘ and ‗compost tea‘ are colloquial and ill-defined terms,

although there have been attempts to better define the term (e.g. Ingham, 1999b).

The dark brown coloured solution that leaches from a pile of compost may be

defined as compost leachate, while the solution obtained after soaking a compost

(either as a slurry or contained in a sack, the latter to remove particulates that

12
Chapter 1: Literature review

would foul irrigation or misting systems if to be used in that manner) in water,

typically for 2 to 21 days (e.g. Weltzein, 1989), may be defined as compost

extract (Diver, 2002). A water-based preparation of compost, incubated with

various microbial food sources (such as molasses, fish and kelp hydrolysate), and

in some cases inoculated with specific microbiota, is best defined as a

‗microbially enhanced compost extract‘.

Compost extracts contain living biota, insoluble organic matter, soluble

nutrients (such as ammonium and phosphate) and organic compounds, such as

proteins, amino acids, carbohydrates, sugars, plant growth regulatory compounds,

siderophores, tannins, phenols and humic acids (Antonio et al., 2008; Dianez et

al., 2006; Ingham, 2000). These components can stimulate plant growth and

suppress soil-borne pathogens (Ingham, 2005; Joshi et al., 2009; Litterick et al.,

2004; Scheuerell & Mahaffee, 2004).

5.2. Production methods of compost extract

Principal production variables include choices of inoculum (compost

feedstock, age), inoculum:water ratio, water quality, added nutrients, incubation

time, solution oxygen level, temperature and pH (Alms, 2002).

Self-evidently, the microbial population of a compost extract will be

dependent on that of the inoculant compost. This consideration extends to the age

of the compost.

Compost extract can be produced by direct addition of a source of

inoculum, typically compost, to water amended with microbial food sources.

However, the resultant product will require filtering to remove particulate matter

to prevent clogging of emitters or spray nozzles (Weltzien, 1991). This issue can

13
Chapter 1: Literature review

be avoided by containing the compost in a permeable bag (e.g. made of muslin,

cotton cloth, nylon stockings or similar) (Ingham, 2005). Following incubation,

the solid materials of the remaining compost can be returned to the composting

pile (Campbell, 2006) or used as a soil amendment.

The compost to water ratio (w/w) utilised in published scientific studies

involving compost extract ranges from 1:1 (Zhang et al., 1998) to 1:50 (Weltzien,

1990). Many studies have followed the method developed by Weltzien‘s

laboratory, employing 1:3 to 1:10 ratios (Haggag & Saber, 2007; Scheuerell &

Mahaffee, 2004; Welke, 2000). Given the high concentration of inoculum to

water, this method may be described more in terms of microbial extraction rather

than microbial population growth. Commercial production of compost extract

may involve considerably lower compost to water ratios (e.g. SmartBugs Pty Ltd,

Lismore, Australia) recommends use of their LivingSoilTM inoculum at a rate of

10 g/L or 1:100). If the compost:water ratio is too low, there is a risk that

contaminating organisms may overwhelm the culture. There is little published

work on the effect of dilution ratio on plant health. For example, in a container

medium study, Scheuerell and Mahaffee (2004) observed significantly reduced

suppression of seedling damping-off disease caused by Pythium ultimum at

higher dilutions (1:9, 1: 4, 1: 1) compared to the neat compost extract, i.e. at 1: 0

dilution. The effect of dilution ratio on the community of extracted and cultured

microorganisms should be studied further.

Typical additions made with the intent of acting as microbial food source

include a carbohydrate source (e.g. molasses) and a N source (e.g. baker‘s yeast,

sea kelp or fish hydrolysate) (e.g. Ingham, 1998). These materials are typically

added during incubation (to support microbial population growth in the culture).

14
Chapter 1: Literature review

Similar additions may be made just before application to the crop, or as a

concomitant application to the crop, to support microbial growth in the crop

environment (Scheuerell & Mahaffee, 2002). Molasses is commonly used in

Queensland and northern NSW where this carbohydrate source is widely

available and rather economical. Generally, molasses is added at a concentration

of 0.5 to 2% to a compost extract batch. Elsewhere, other sugar sources may be

used, for example, Sackenheim et al. (1994) reported use of 3% sucrose (in both

aerated and non-aerated compost extracts).

Further, additional inorganic and organic additives, such as humic acid

and rock dust, may be included with the aim of enhancing microbial growth. For

example, addition of rock dust and humic acid to aerated compost extract has

been reported to enhance the effectiveness of the compost extract in managing

plant disease caused by Botrytis cinerea (Scheuerell & Mahaffee, 2006).

A typical guide to the quality of water required for compost extract

production is for water fit to drink, with rain water commonly used. Use of

surface waters (e.g. dam water) risks biological contamination of the compost

extract. Use of sterilised surface water or town water supply risks chemical

contamination of the compost extract. Where chlorine has been used for

sterilisation, vigorous stirring for about half an hour is recommended to de-gas

chlorine from water (Ingham, 2003).

The pH of the culture solution will also affect the growth and diversity of

microorganisms, with bacterial population growth favoured at neutral pH while

fungal (including yeasts) growth is favoured under acidic and alkaline conditions

(Ingham, 2003). In a large scale food composting, microbial activity as indicated

by the carbon-di-oxide production was inhibited by low pH below 6 during the

15
Chapter 1: Literature review

early phase of composting (Sundberg et al., 2004). Rousk et al. (2010) found that

pH has a significant influence on the phospholipid fatty acid compostion of

microbial community as the community changed along the pH gradient of soil,

showing a correlation of 0.97.

The microbial population composition of compost extract can also be

expected to be affected by the incubation temperature. Scientific publications on

the impact of variable temperature on microbial population and composition of

soil as well as on composting process are numerous (Feng & Simpson, 2009;

Herrmann & Shann, 1997; Holtan-Hartwig et al., 2002; Tang et al., 2007; Uvarov

et al., 2006). For example, Tang et al. (2007) found that the microbial community

pattern (DGGE) during the thermophilic and mesophilic composting processes

changed at different incubation temperatures of 60˚C and 30˚C. The authors

found a relatively lower population and diversity of microorganism, higher

biomass of Proteobacteria and lower biomass of Actinobacteria in the mesophilic

compared to the thermophilic compost.

Parallel studies on the effect of temperature on the microbiology of

compost extracts should be undertaken.

Compost extract can be produced under aerobic or anaerobic conditions,

with obvious consequences to the microbial community supported. The

anaerobic, or fermentation, method involves incubation of compost in water with

or without amendment and without extra supply of aeration (Diver, 2002),

producing non-aerated compost extract. Even without enclosure of the container,

anaerobic conditions will prevail through most of the vessel given the slow rate

of diffusion of oxygen through water. Fermentation for up to several weeks, with

occasional stirring, is practiced. Non-aerated compost extract is typically

16
Chapter 1: Literature review

produced using longer fermentation times than that used for incubation of aerated

compost extract, to allow for greater extraction (separation of substances from the

solid matrix to liquid) of nutrients and beneficial microbes.

This production of non-aerated compost extract is simpler than that of

aerated compost extract (avoiding the need to maintain aerobic conditions); but

the reduced oxygen condition will lead to nutrient losses, particularly through

volatilisation of N following denitrification. Further, the community of obligate

and facultative anaerobes produced in such a culture may have little relevance to

that of the aerobic soil. Also, phytotoxic materials, such as organic acids, phenols

and ammonia may be produced by the anaerobic microorganisms (Campbell,

2006; Ingham, 2005).

To produce aerated compost extract, the extract is aerated to maintain a

level of dissolved oxygen not less than 6 ppm (Ingham, 2005). The duration of

incubation is usually 2-3 days. Ingham (2000) reported that the optimum

incubation time should be characterised in terms of when the maximum active

microbial biomass is achieved, and indicated that this commonly occurs in 18-24

hours with commercial extract-making equipment and procedures.

As a batch culture, respiratory demand is not constant over the incubation

period. At peak respiratory demand, there is a risk of decreased oxygen levels in

the solution. Some operators (O‘Grady, 2008; pers. comm.) suggest that short

periods (presumably a few hours) of dissolved oxygen less than 6 ppm are not

detrimental to the development of the microbial community.

Aerobic incubators have become commercially available in recent years.

The basic components include the container (typically 60 to 5,000 L), fitted with

a drainage point that can be connected to a dispensing pump, an aeration system

17
Chapter 1: Literature review

serviced by an air blower, and an optional compost basket (Diver, 2002). Earlier

designs involved circulation of the extraction solution, with cascade systems

aiming to provide solution aeration, or circulation over the compost with the aim

of improving extraction. However, at present, improved designs to meet the high

oxygen demand of the culture are available in the market (e.g.

www.smartbugs.com.au, www.soilsoup.com, www.greenorganics.biz,

www.trustnature.com.au, www.composttea.com, www.cleanairgardening.com,

www.jollyfarmer.com). In general, current systems employ a much higher rate of

gas flow to achieve aeration, and this process serves the secondary purpose of

vigorously mixing the solution.

The rate of oxygen diffusion through water is 10,000 times less than that

through air. Efficient aeration systems will therefore maximise the surface area of

the gas-solution interface. This requirement drives the use of aeration systems

that produce smaller bubbles and that maximise the residency time of the bubble

in the solution column. Aquarium stones, as recommended, for example, by

Diver (2002), are typically inadequate to maintain the required rate of

oxygenation. Various aeration mechanisms have been recommended, for

example, vortex-mixing high pressure nozzle systems for entrainment of air into

the water stream (Ingham, 1999a), membrane diffusers (smartbugs.com.au) or

micro incubators fitted with a pump and hose (www.asapsupplier.com) to infuse

air into the compost extracts.

The air pumps required for these systems are characterised by a low head,

high volume specification. Few compost extract incubators would involve a

solution depth of more than 2 m (static pressure of 0.2 atm), and generally less

than 1 m. However, the aeration (bubbling) system may have a high resistance to

18
Chapter 1: Literature review

gas movement, necessitating use of a pump capable of delivering a higher

pressure.

In practice, the aerator is stopped for about half an hour to allow insoluble

materials to settle prior to using the extract (Campbell, 2006). At this point the

batch culture may enter a plateau phase in population growth. Population growth,

however, can be restarted by addition of food sources.

In the plateau phase, the culture maintains an (albeit lower) respiration

rate. Storage of this material without aeration therefore risks development of

anaerobic conditions and change in the species composition of the microbial

culture. It is therefore generally recommended that the compost extract be applied

as soon as possible after production, and if stored it should be in a well-ventilated

tank with adequate agitation (Bess, 2000). However, some users (e.g. O‘Grady,

2009; pers. comm.) suggest that amending the cultures to low or high pH might

cause the microbiota to enter a dormant phase, which may allow for storage or

long distance transport of the material. Soil Foodweb Institute

(www.soilfoodweb.com.au) suggests that such ‗put-to-sleep‘ extracts can also be

produced by prolonging incubation time or reducing the activity of microbiota

with some chemicals, however there are chances of losing around 50% of the

microbial diversity in this process.

Cleaning of incubators after incubation of compost extracts is important,

as residue left inside the unit may generate anaerobic conditions and favour

contaminating organisms such as Escherichia coli, Pseudomonas spp. (Alms,

2002). Hypochlorite, hydrogen peroxide and detergents can be used effectively to

clean the inside of the incubator (Ryan, 2003).

19
Chapter 1: Literature review

5.3 Defining compost extract quality

As noted previously, compost extract quality is dependent on the nature

and content of inoculum (compost) used (Carpenter, 2005), and also to conditions

prevailing during the incubation process. Unfortunately, there can be

considerable batch-to-batch inconsistency in compost extract composition

(Campbell, 2006; Ingham, 1999a). In general, a compost extract that has high

level of plant available nutrients and high species richness, evenness and total cell

number of beneficial microorganisms is desired, that is, when used to amend soil,

the compost extract should have organisms that play key functions in the soil.

While there is no officially recognised standard on what constitutes a

good quality compost extract, either in Australia or elsewhere, ‗private label‘

quality standards do exist.

In Australia, the Soil Foodweb Institute (Lismore, NSW;

www.soilfoodweb.com.au) offers a service in (biological) quality assessment of

compost extract and soils using both qualitative and quantitative methods.

Ingham (2005) has reported that compost extract can be qualitatively assessed in

terms of microbiota diversity and density by microscopic examination. A whole

mount preparation of compost extract is examined at 400X magnification, and

biota in up to ten fields of view assessed. The compost extract is assessed as poor

if no or very few microbes are present, good if more than 500 bacteria and at least

one fungal hypha are present in five fields, and very good if innumerable bacteria

and at least one fungal hypha are present in first five fields. If innumerable

bacteria and at least one fungal hypha, of rich species diversity, and high numbers

of protozoa (flagellates, amoeba and ciliates) and nematodes are present, the

compost extract is assessed as excellent. Soil Foodweb Institute (SFI) also offers

20
Chapter 1: Literature review

a quantitative assessment service. Viable cell counts are presumably established

using visualisation by epi-flourescence microscopy, with fluorescein diacetate

staining (e.g. Ingham, 2004). Quantitative assessment requires the assessment of

a known volume of sample and the bacterial and fungal biomass is presumably

estimated based on a calculation involving cell numbers and dimensions. SFI

measures a value of plant available N, presumably from the microbial biomass

figures. SFI also measures nematode density and species diversity to genus level,

as this is considered to index soil ecosystem health.

Quantitative analysis of soil microbiota following the application of

compost extracts at field level to index soil health is crucial to demonstrate any

effect of compost extract use.

6. Characterising microbial diversity

A range of techniques are available for the study of microbial diversity in

environmental samples. Perhaps the most commonly used of these techniques is

simple microscopic examination; the plate count method and the fluorescein

diacetate hydrolysis method are also in use at present. However, these traditional

methodologies may assess only a small fraction of the total microbial diversity

(Scheuerell, 2004). Such limitations of culture-based techniques underestimate

the total representation of microbiota because the microorganisms may remain in

dormant but viable states in the environmental samples but usually fail to grow

on Petri plates. Complex inter-dependencies between microorganisms in the

environmental samples cannot be generated in the laboratory. DNA based

methodologies, as discussed later, may offer potential for greater insight into the

microbial community (diversity) of compost extract.

21
Chapter 1: Literature review

The plate count method is a standard method for microbial enumeration,

generally expressed in colony forming units (cfu). This method can either be

spread plate or pour plate or by placing a membrane filter, after filtering the

tested material, on the plate surface. The preferred method is to serially dilute the

sample and to place the aliquot (usually 1 mL) onto the surface of the plates or to

pour the aliquot to the agar media to determine the cfu mL-1 (g) of the initial

sample.

The fluorescein diacetate (FDA) hydrolysis method is a simple and widely

used spectrophotometric method in which the enzymatic activity of

microorganisms is measured. This approach provides an estimate of total

microbial activity (Adam & Duncan, 2001; Green et al., 2006; Schnurer &

Rosswall, 1982). The assay is sensitive to the activity of several ubiquitous

enzymes, such as lipase, esterase and protease, and thus is considered as non-

specific. Activity of these enzymes results in the hydrolysis of FDA releasing

fluorescein (Fig. 1.1) which can be measured spectrophotometrically.

Figure 1. 1. Enzymatic conversion of FDA to fluorescein (Source: Adam & Duncan, 2001)

This technique has been widely used with environmental samples, for

example, soil (Schnurer & Rosswall, 1982); activated sludge (Fontvieille et al.,

1992); river sediments (Marmonier et al., 1995); and in compost (Ntougias et al.,

2006). This technique should therefore be applicable to the estimation of

22
Chapter 1: Literature review

microbial activity of compost extract although there are other techniques

available for the measurement of enzymatic (e.g. dehydrogenase, urease,

phosphatase) and biochemical (e.g. ergosterol) activities of various environmental

samples.

Techniques based on microbial culture on a range of food sources have

also received great attention in recent years to characterise microbial community

diversities from a range of ecological samples. BiologTM Inc, USA offers a wide

range of single C sources supplied in 96 well plates, and based on their utilisation

pattern, microbial functional diversity is estimated (Preston-Mafham et al., 2002).

To measure the community level physiological profile (CLPP), which is

the metabolic fingerprinting of the microbial community, Biolog Ecoplate and FF

microplate are common in use. The Biolog Ecoplate has 96 wells consisting of 31

single C sources and the control, replicated three times, in addition to a

tetrazolium dye. The Biolog Ecoplate is specially designed for measuring

functional diversity of the bacterial community. The 31 C sources are divided

into six groups of carbohydrates (7), carboxylic acid (9), polymers (4), amino

acids (6), amines (2) and miscellaneous substrates (3). The utilisation of any C

source by the bacterial community results in the respiration dependent reduction

of tetrazolium dye and purple colour formation that can then be quantified and

monitored over time.

Similarly, the FF microplate is designed for measuring fungal functional

diversity utilising 95 single C sources, which were divided into six groups of

carbohydrates (44), carboxylic acid (17), polymers (5), amino acids (13), amines

(6) miscellaneous substrates and (10) the control (without any C sources). Here,

iodonitrotetrazolium violet is used as a redox dye to colorimetrically indicate the

23
Chapter 1: Literature review

mitochondrial activity during the oxidation of C sources producing reddish

orange colour (Insam, 1997). The carbon substrates found in the Ecoplate and FF

microplate from BiologTM, respectively, divided into different substrate guilds

according to Preston-Mafham et al. (2002), are presented in Table 1.1.

24
Chapter 1: Literature review

Table 1. 1. Carbon substrates of Ecoplate and FF microplate (BiologTM), respectively divided into different substrate guilds according to Preston-Mafham et al. (2002).

Carbohydrates Carboxylic acids Polymers Amino acids Amines/amides Miscellaneous


-Methyl-DGlucoside D-Galactonic Acid ã-Lactone Tween 40 L-Arginine Phenylethylamine Pyruvic Acid Methyl Ester
D-Xylose Dgalacturonic Acid Tween 80 L-Asparagine Putrescine Glucose-1-Phosphate
i-Erythritol 2-Hydroxy Benzoic Acid  -Cyclodextrin L Phenylalanine D,L--Glycerol Phosphate
D-Mannitol 4-Hydroxy Benzoic Acid Glycogen L-Serine
N-Acetyl-DGlucosamine -Hydroxybutyric Acid L-Threonine
D-Cellobiose D Glucosaminic Acid Glycyl-Lglutamic Acid
 -D-Lactose Itaconic Acid
 -Ketobutyric Acid
D-Malic Acid
N-Acetyl-DGalactosamine D-Mannose D-Galacturonic Acid Tween 80 -Amino-butyric Acid D-Glucosamine Amygdalin
N-Acetyl-DGlucosamine D-Melezitose D-Gluconic Acid  -Cyclodextrin L-Alanine Glucuronamide Glucose-1-Phosphate
N-Acetyl-DMannosamine D-Melibiose D-Glucuronic Acid -Cyclodextrin L-Alanyl-Glycine Succinamic Acid Glycerol
Adonitol  -Methyl-DGalactoside 2-Keto-D-Gluconic Acid Dextrin L-Asparagine Alaninamide Salicin
D-Arabinose -Methyl-DGalactoside Fumaric Acid Glycogen L-Aspartic Acid 2-Amino Ethanol Bromosuccinic Acid
L-Arabinose  -Methyl-DGlucoside -Hydroxy-butyric Acid L-Glutamic Acid Putrescine D-Lactic Acid Methyl Ester
D-Arabitol -Methyl-DGlucoside -Hydroxy-butyric Acid Glycyl-L-Glutamic Acid Succinic Acid Mono-Methyl Ester
Arbutin Palatinose p-Hydroxyphenylacetic Acid L-Ornithine Adenosine
D-Cellobiose D-Psicose  -Keto-glutaric Acid L-Phenylalanine Uridine
i-Erythritol D-Raffinose L-Lactic Acid L-Proline Adenosine-5'-Monophosphate
D-Fructose L-Rhamnose D-Malic Acid L-Pyroglutamic Acid
L-Fucose D-Ribose L-Malic Acid L-Serine
D-Galactose Sedoheptulosan Quinic Acid L-Threonine
Gentiobiose D-Sorbitol D-Saccharic Acid
-D-Glucose L-Sorbose Sebacic Acid
m-Inositol Stachyose Succinic Acid
-D-Lactose Sucrose N-Acetly-Lglutamic Acid
Lactulose D-Tagatose
Maltitol D-Trehalose
Maltose Turanose
Maltotriose Xylitol
D-Mannitol D-Xylose

25
Chapter 1: Literature review

Serially diluted environmental samples are incubated to each well of the

Biolog Ecoplates and FF microplates for a few days and optical densities are

measured at 24 h intervals to obtain the kinetics of average well colour

development (AWCD) of bacterial and fungal communities. The AWCD and

optical density (OD) are calculated for all different C substrates utilised in

Ecoplate and FF microplates before statistically analysing the data (Garland,

1997).

DNA-based methods have enormous potential to underpin knowledge of

structural and functional characteristics of microbial communities. Such methods

have been employed widely within the field of microbial ecology to analyse

diversity and composition of microbe communities. Most of these methods rely

upon the use of polymerase chain reaction (PCR) using group specific primers

associated with selected ribosomal DNA (rDNA) regions followed by digestion

with endonuclease restriction enzymes and sequencing of selected ribosomal

DNA (rDNA) regions (Dees & Ghiorse, 2001; Ntougias et al., 2004). These steps

are followed in amplified ribosomal DNA restriction analysis (ARDRA)

(Koschinsky et al., 1999), a commonly used fingerprinting method, which allows

grouping of the organisms under study by simple restriction endonuclease

digestion of PCR products.

Denaturing gradient gel electrophoresis (DGGE) (Muyzer et al., 1993) or

temperature gradient gel electrophoresis (TGGE) (Muyzer, 1999) are also used in

this context. The following steps are as for ARDRA, with cloning and sequencing

allowing phylogenetic analysis. The phylogenetic affiliation of the predominant

community members, as represented in the DGGE band pattern, can be inferred

by comparative analysis of sequences from excised and re-amplified DNA

26
Chapter 1: Literature review

fragments (lanes 2-7 as shown in Figure 1.2) and sequences stored in nucleotide

databases (Muyzer, 1999).

To investigate the genetic diversity of the microbial populations (at

species level) in the environmental samples, DGGE followed by cloning,

sequencing and phylogenetic analysis of 16S rDNA is the most commonly used

approach (Fig. 1.2). DGGE has been used mostly in sediments and soils for

complexity and community change studies (Muyzer & Smalla, 1998; Postma et

al., 2008). DGGE has generally been used to analyse the genetic diversity of

bacterial populations in environmental samples (Muyzer et al., 1993), but the

technique has also been applied to fungal communities (Anderson, 2003; Jany &

Barbier, 2008). While there are some studies on microbial diversity in compost

using DGGE (Adams & Frostick, 2009), investigation of diversity in compost

extracts, and the impact of extracts on soil microbial diversity, is warranted.

27
Chapter 1: Literature review

Figure 1. 2. Flow diagram showing the different steps in the analysis of microbial community
structure by PCR-DGGE. DNA is extracted from an environmental sample and used as a template
to amplify (PCR) the 16s rRNA encoding genes of microorganisms (e.g. bacteria). Thereafter, the
PCR products are separated by denaturing gradient gel electrophoresis (DGGE), which is
followed by cloning, sequencing and phylogenetic analysis of 16S rDNA.

For DGGE analysis, total DNA extraction is followed by polymerase

chain reactions (PCR) using universal primers and a thermocycler (Girvan et al.,

2003). The PCR-amplicons are subjected to agarose gel electrophoresis and

visualised after staining under UV irradiation. Following electrophoresis, the

DGGE gels are silver stained (Girvan et al., 2003) and scanned using advanced

analysis package. The UPGMA (Unweighted Pair Group Method with Arithmetic

mean algorithm) dendrogram is generated and the digitised image is analysed to

express the relatedness of microbial communities as similarity clusters. The

Shannon-Weaver diversity index, H‘ (Shannon & Weaver, 1963) and Equitability

index, J (Begon et al., 1990) are calculated from the number and intensities of

bands present in each lane with the following equations: H‘ = -∑(piLNpi), where,

pi = ni/N, ni is the intensity of the band i, N is the total intensity of all bands in

28
Chapter 1: Literature review

the lane and LNpi is the natural log of pi; J = H‘/(LNni), where H‘ is the

Shannon-Weaver diversity index and LNni is the natural log of the total number

of species in a lane. The Shannon-Weaver diversity index is thus a measure of

biodiversity which takes account of the number and evenness of species in the

community profile, while the Equitability index rates only the overall evenness of

band intensities and the S value the number of bands per sample (Dilly et al.,

2004; Krebs, 1999). The DGGE generated bands can be excised from the gels,

sequenced, and tested up to species level to determine their phylogenetic

associations with the database (e.g. www.ncbi.nlm.nih.gov) of GeneBank (Jing et

al., 2010).

7. Current use of compost extract in central and southern Qld

and northern NSW

The use of compost extract is accepted within the organic farming

community, and, in something of a social phenomenon, is gaining acceptance

amongst conventional broadacre farmers and some graziers in central and

southern Qld and in northern NSW. This trend is prompted by farmer interest in

soil health and by the rising price of fertilisers. Users report that they can cut

costs without compromising yield. It appears that a dramatic rise in the rate of

uptake of this technique will occur over the next few years.

Users of compost extract in central Queensland and northern NSW

include certified organic horticultural small crop producers on the Capricorn

coast (generally small scale), broadacre wheat/sorghum/chickpea and cotton

producers of the central highlands, and graziers (Table 1.2).

29
Chapter 1: Literature review

Table 1. 2. List of farmers [permissions granted to use their data] from Qld and NSW visited with Tom Nicholas who are using compost extracts in different crops, and their
claimed benefits (2008).

Area of Rate of Method of Frequency of


Farmer/Contact Crops production application Dilution application application/year Claimed benefits
Lex Webb, Baralaba, QLD Wheat in rotation to 2500 ha Compost extract - Foliar spray and - Improves growth and yield
Tel: 0428581162 sorghum / mungbean @ 30 L/ ha liquid injection
E-mail: belvedere@bluemax.com.au CalSap @ 3L/ha

Paul Murphy, Capella, QLD Spelt wheat and 900 ha Compost extract 20% water addtion Seed priming Once Improves root growth, soil carbon
Tel: 049816767 / 0428816768 bread wheat @ 8 L /ton of seeds to the concentrate and microbial population
E-mail: murphyfarming@bigpond.com CalSap @ 2L/ha compost extract
from CQ Compost

Bryson Taylor, Clermont, QLD Sorghum, chickpea, 2200 ha Compost extract Concentrate compost Liquid injection Once -
Tel: 0749835365 / 0427125022 wheat @ 30L/ha extract obtained
E-mail: kilmacolm@bigpond.com from CQ Compost

Dave Daniels, Clermont, QLD Wheat and chickpea 2830 ha Compost extract Concentrate compost Seed priming Once Improves soil biology, release of
Tel: 0749835272 @ 30 L/ha extract obtained locked-up nutrients, improves soil
E-mail: daniels.travellon@bigpond.com UAN @ 20L/ha from CQ Compost structure, increases soil carbon,
CalSap @ 3L/ha better crop

Tom Nicholas, Clermont, QLD Fallow/field covered 2000 ha Compost extract - Liquid injection Once Better root system, grain quality
Tel: 0749835252 / 0428835463 with sorghum stubble @27 L/ha in seed furrow and better growth of plants
E-mail: tgnicholas@gmail.com CalSap @ 3L/ha

Stuart Larson, Mallanganee, NSW Cereals, soybean, 2040 ha Compost extract - Foliar spray Once Improved soil moisture
Tel: 61266645145 / 0427645171 barley, pasture @ 190 L/ha and organic matter,
E-mail: stuartlarsson@maraseeds.com.au cheap source of C,
cost effective

Geoff Brown, Spring Ridge, NSW Wheat, barley 1080 ha Compost extract 1:50 (w/v) Foliar spray Twice or Insect and disease control,
Tel: 0267473848 / 0428456088 @ 20-40 L/ha more low input cost, improvement
E-mail: goodgerwirri@activ8.bnet.au Molasses, fish emulsion, in growth and yield of crops,
liquid kelp and humic no residual chemicals
acid @ 2 L each/ha

Colin Seis, Winona, Gulgong, NSW Oats, cereal rye 840 ha Compost extract 1:500 (w/v) Liquid injection Once Yield comparable to chemical
Tel: 0263759256 @ 50 L/ha fertiliser, no incidence of
E-mail: colin@winona.net.au Molasses, fish emulsion, disease recorded in the last ten
liquid kelp and humic years, cost effective
acid @ 2 L each/ha

30
Chapter 1: Literature review

The increase in use of compost extracts is being supported by a range of

commercial suppliers of either soil biological amendments including compost

extracts, compost extract incubating tank designers and compost extract and soil

biological quality assessment services (Table 1.3). For example, CQ Compost Pty

Ltd (Emerald) is a major supplier of compost extracts and relevant amendments,

primarily to broadacre cotton/wheat/sorghum/chickpea growers. Compost

extracts are distributed in large volume tanks (1000 L) to farmers. BioNutrient

Solutions Pty Ltd, Moree, NSW was a previous producer of compost extract, but

ceased production because of difficulty in managing the liquid extracts (B.

Davidson, 2008; pers. comm.).

31
Chapter 1: Literature review

Table 1. 3. List of companies [permissions granted to use their data] visited with Tom Nicholas or contacted that were involved in producing compost extract and relevant
organic and biological products and/or providing services and their claims (2008).

Company/Contact Products/Services Claims

CQ Compost, Emerald, QLD Compost extract from compost made from Improves soil health
Tel: 0749876511 / 0427876643 wormcast
Earthlife Pty Ltd., Toowoomba, QLD Several agricultural products for garden, Increases soil organic matter, and soil carbon,
Tel: 07 4633 2219 horticulture and broadacre improves crop quality
E-mail: enquiries@earthlife.com.au
Web: http://www.earthlife.com.au
McLeod's Agriculture, Toowoomba, QLD Organic soil conditioner, zeolite, soil microbes, Improves soil aggregation, organic matter,
Tel: 1800 062616, 07 46993360/0419669115 liquid fertilisers nutrition, cut costs, adds biology in soils
Fax: 0746993359
E-mail: mcleodsagriculture@yahoo.com.au
Web: http://www.mcleodsorganicfertiliser.com/
Grazing BestPrac, Yeppoon, QLD Education and extension service aimed at Best- Encourages regenerative grazing practices to
Tel: 07 4938 3919 Practice for grazing - primary services include improve soil carbon and to improve soil health
E-mail: mick@grazingbestprac.com.au farmer training workshops and consulting by microbial management, produces compost
Website: www.grazingbestprac.com.au focussed on grazing management and strategies extracts and molasses extracts, develops
to develop healthy soils sustainable strategies and increases profits
BioNutrient Solutions Pty Ltd., Moree, NSW Bionutrient solutions, fertilisers, seed treatments, Improves soil aggregation, nutrition, organic
Tel: 0732711058 liquid humate and many other products matter, cut costs, adds biology in soils
E-mail: bart@bionutrient.com.au
Web: http://www.bionutrient.com.au

Soil Food Web Institute, Lismore, NSW Quantitative assessment for soil, compost and Suppresses foliar and root diseases, improves
Tel: 0266225150 compost tea (extracts) measuring biomass by plant nutrition, health, and crop yield
E-mail: contact@soilfoodweb.com.au direct count methods: bacteria, fungi, protozoa,
Web: www.soilfoodweb.com.au nematodes, mycorrhizal root colonization,
consultancy service
O'Grady Rural, South Lismore NSW Regenerative agricultural consultancy/education, Encourages regenerative agricultural practices to
Tel: 02 66216088 / 0429200492 produces and markets microbial nutrients and improve soil carbon and improve soil health
Email: rogrady@bigpond.net.au microbial fermentation equipment (e.g. compost
Web: www.smartbugs.com.au tea brewers), biochar and pandachar, conducts
farm-ready workshops to assist farmers achieve
these outcomes

32
Chapter 1: Literature review

8. Compost extract application and claimed benefits

Users and producers of compost extract have claimed benefits (e.g. Alms,

2002; Elad et al., 1996) in terms of enrichment of soil nutrient level and

enhancement of populations of specific soil microorganisms. Microbial

enhancement is claimed to provide nutrient turnover in the soil, suppress

diseases, remediate soil contaminants, increase soil C levels particularly through

accumulation of humates, improve soil physical structure, improve water holding

capacity and to increase anion/cation exchange capacity (Adams, 1990; Bess,

2000; Cook & Baker, 1983; Griffiths, 1965; Scheuerell & Mahaffee, 2002;

Touart, 2000).

Compost extract is commonly applied to the soil with the aim of

enhancing soil health. Application methods include soil drenching, targeting the

root zone (rhizosphere) of the plants, for example, using planters modified to

deliver a stream of compost extract into the planting furrow, and via irrigation

systems to established plants (Sebti, 2005). Soil application may be undertaken

with a view to suppress disease and/or to improve plant nutrition (Litterick et al.,

2004). Soil application of compost extract (10 L tree-1) plus four foliar

application of yeast (1%) and humic acid (0.5%) is reported to increase fruit yield

and quality with improvement in nutrition status of olive trees compared to the

control treatment (Fayed, 2010b).

Direct application of compost extract to plant foliage is also reported

(Elad & Steinberg, 1994; Ryan et al., 2005; Segarra et al., 2009; Weltzien, 1991).

Spraying is done with a broad range of sprayers, however, mist spraying is the

preferred method (Grobe, 2003). Foliar application is intended to alter the

composition of the microbial population on the leaf surface through nutrient

33
Chapter 1: Literature review

supplementation (Fayed, 2010a) and direct addition of microbes, and is generally

undertaken with a view to suppress disease (Ingham, 1999a). Application under

moist conditions has obvious relevance to microbial survival. Also, it is

recommended that compost extract be applied either in the morning or evening on

sunny days because the microorganisms are sensitive to UV light (Ingham, 2005).

The rate of compost extract application varies significantly according to

intended aim and mode of application (Ingham, 2005). For foliar application,

reported application ranges from 20-50 litres per hectare per application

(sufficient to wet both surfaces of all leaves). This further depends on the canopy

area of crops (Campbell, 2006). Four to six foliar applications may be made to a

crop during its growth period, however, frequency may increase with disease

infestation. S. Stevens (2008; pers. comm.) suggests that soil application of

between 150-200 litres per hectare per application is typical but lesser amounts

are used when compost extract is applied within the seed furrow. However, it is

recommended to apply sufficient volume of compost extract so that it reaches the

entire root zone (Scheuerell, 2003). Variables in application procedure include

the application equipment, dilution ratio, and use of spray adjuvant and the

concomitant application of a microbial food source.

The fundamental aim of compost extract incubation is to alter microbial

populations towards a more plant-beneficial community (Hoitink & Boehm,

1999a). The working principle of compost extract varies from that of the

synthetic chemicals in the way that compost extract adds microorganisms to the

soil and ‗brings the soil back to life‘ (Martens, 2001). There are also a range of

other biological amendments (e.g. direct compost addition, incorporation of green

34
Chapter 1: Literature review

manure crops) used with the aim of restoring and regenerating soil microbial

communities (Larkin, 2008).

Examples of specific claims include use for fertility management

(Scheuerell & Mahaffee, 2002) as compost extract is considered to add nutrients

(Reeve et al., 2010), to improve soil structure by building soil aggregates and to

increase water holding capacity of soils (Ingham, 2005). There is also a growing

body of published work on the efficacy of compost extract in plant disease

suppression (Brinton, 1995; Kai et al., 1990; Koné et al., 2010; Siddiqui et al.,

2008a; Zmora-Nahum et al., 2008).

The hypothesised benefits claimed by proponents of compost extract can

be categorised as follows:

 direct nutrient addition

 soil inorganic fraction solubilisation

 soil organic matter turnover (mineralisation)

 leaf litter and stubble decay

 increased free living N2 fixation

 suppression of plant diseases

 acquired systemic resistance

 bioremediation

 improved soil structure

 improved water holding capacity and

 improved soil C content.

35
Chapter 1: Literature review

8.1 Direct nutrient addition

Compost extract contains a range of soluble elements essential to plant

growth and yield enhancement (Fayed, 2010a; Ryan, 2003; Sebti, 2005; Siddiqui

et al., 2008b). This source will be significant to plant growth when compost

extract is applied at high rates, but is generally insignificant when compost

extract is applied in broadacre applications. When used in broadacre applications,

compost extract is generally used in ‗spot‘ applications, for example, when

applied with seed in the plant furrow. As such, it may have localised nutritive

value (e.g. supporting early seedling establishment).

Some examples of direct influence of compost extracts have been

demonstrated in some studies. For instance, the growth and performance of pak

choi plants were improved when compost extracts were applied as a soil drench

or foliar spray (Pant et al., 2009). The authors reported an improvement in an

uptake of both macro and micro nutrients (e.g. total N, P, K, Ca, Mg and S) as

well as greater total carotenoid content. Likewise, Sebti (2005) demonstrated

increased growth, yield and biomass of a tomato crop compared to the water

control and found that growth effects of compost extracts were comparable to

those of compost and organic fertiliser (Guanito) when applied through foliage or

through fertigation. Soil and foliar applications of compost extracts on

pomegranate trees at the rate of 5 L per tree combined with an antioxidant spray

of ascorbic acid and citric acid mix produced the highest yield with better fruit

quality as indexed by greater vitamin C, total sugar, total soluble solid and total

anthocyanin compared to that of other treatments.

36
Chapter 1: Literature review

8.2 Soil inorganic fraction solubilisation

A range of plant essential elements are present in insoluble inorganic

forms in the soil. Soil microbiota may assist in accessing this resource, effecting a

transfer from the insoluble to the soluble. Perhaps the best example of this

function is for phosphorus (P). Relatively large amounts of P can be locked up in

insoluble forms in the soil (e.g. calcium phosphate) limiting plant growth

(Raghothama, 1999). Indeed much applied (inorganic or organic) P is ‗lost‘ to

insoluble forms. Phosphate solubilising bacteria, such as Pseudomonas, Bacillus

and Rhizobium spp. and fungi, especially mycorrhizal fungi, can mobilise this

source (Rodríguez & Fraga, 1999). For instance, rhizosphere microorganisms

associated with phosphate solubilisation have been reported in the mangroves of

a semiarid coastal lagoon (Vásquez et al., 2000). Furthermore, the reported

growth stimulating action of several isolates of Trichoderma on beans has been

attributed to their roles in producing auxins, siderophores or by phosphate

solubilisation in the rhizosphere (Hoyos-Carvajal et al., 2009). The possible

mechanisms for converting phosphate into a soluble form involve acid

production, redox activity or chelation by metabolites released into the

rhizosphere (Altomare et al., 1999).

Similarly, strains of fluorescent Pseudomonas promoted growth of

various crops, such as rice, maize and sorghum, and demonstrated their potential

to be used as bioinoculants or biofertilisers (Suresh et al., 2010). These strains

were demonstrated to possess antifungal properties and to produce siderophores,

hydrogen cyanide, proteases and plant growth hormones, such as indole acetic

acid solubilise phosphate. These strains of varying potential for crop benefits are

likely to be present in compost extracts produced from various sources. Janzen et

37
Chapter 1: Literature review

al. (1995) reported that addition of a sterile extract of compost to a phosphate

limiting bacterial growth medium (at 10 ug C mL-1), sourced from

phosphogypsum, which is the residue of rock phosphate obtained during the

manufacture of phosphate fertiliser, increased microbial-P in a culture of

Azospirillum brasilense.

8.3 Soil organic matter decomposition

Soil microbiota play a central role in decomposition of organic residue in

soils, and the rate of this turnover can be increased by microbial enhancement

using compost extract (Ingham, 2005; Ryan, 2003). Some of the fungal species

known to aid in degradation of organic matter, and prevalent in most soils, are

Trichoderma spp. (Haaba et al., 1990) and cellulose digesting brown rot fungi,

such as Coniophora prasinoides, Coniophora puteana (Highley, 1980) and

Cellulomonas spp. (Lines-Kelly, 2004). Naidu et al. (2010) reported the presence

of several advantageous bacteria, such as lactic acid bacteria (production of lactic

acid), Bacillus spp. (biocontrol of disease and N2 fixation), and fungal species,

such as Penicillium spp. (food and drug production), Trichoderma spp. (disease

suppression and leaf liter decay), Aspergillus spp. (medical and commercial

production), yeast (nutritional and health benefits), Streptomyces spp. (disease

suppression through antibiotic production) in compost extracts. These beneficial

microorganisms are generally prevalent in composts and are likely to be water

extracted, with microbial multiplication possible with or without microbial

additives and/or aeration.

The microbial decomposition of soil organic matter can yield humates and

other organic fractions that improve soil physical structure and soil cation/anion

38
Chapter 1: Literature review

exchange capacity (Sparling, 1997). Further, the decomposition-mineralisation

process will make nutrients available to plants (Anderson, 2003; Bourne et al.,

2008).

8.4 Leaf litter and stubble decay

In agricultural production, nutrients are exported from the farm with soil

lost through erosion and in harvest products. For the latter, a wheat yield of 5t ha-
1
year-1 and 1% N will remove 50 kg N ha-1year-1 in grain. Nutrients, such as N,

get lost through gaseous emissions (e.g. nitrous oxide via nitrification and

denitrification), ammonia volatilisation and nitrate ions via leaching.

Soil microbial enhancement can promote the tapping of plant essential

elements from inorganic and organic pools within the soil. However, this equates

to ‗soil mining‘, and such pools will eventually be exhausted unless replenished

by inorganic or organic sources. Admittedly, however, the total reserve in the soil

can be very large for some elements (e.g. phosphorus, potassium, iron), capable

of supporting plant growth for many decades.

The issue of leaf litter and stubble decay is an extension of the soil

organic matter turnover, exacerbated by the dry conditions that often prevail

above-ground. Under dry conditions above-ground dead plant material is

essentially preserved. This is a particular challenge to CQ graziers who would

prefer standing grass biomass converted to soil organic matter rather than lost by

burning (O‘Grady & Alexander, 2008), but also applies to farming practices

involving no or low tillage, and elimination of fire (e.g. of cane harvest trash in

sugar cane production) (Hall et al., 2006). Loss of soil organic matter results in

the loss of microbial biomass, activity and diversity, contributing to soil

39
Chapter 1: Literature review

degradation in sugarcane burning systems prior to harvest (Dominy & Haynes,

2002; Dominy et al., 2002). Graham and Haynes (2005) found lower community

diversity in soils under the pre-harvest burning system compared to that covered

with sugarcane stubble following the green cane harvest. The decline in loss of

microbial diversity can, however, be curtailed either by ceasing burning or by

retaining the trash in the soil (Graham et al., 2002). Microbial diversity could also

be enhanced by applying microbially enhanced compost extracts in the soil over

harvest residues with the aim of accelerating the rate of residue decomposition

(Hall et al., 2006).

8.5 Free living N2 fixation

For N there is an additional potential benefit – that of N2 fixation by

certain bacteria. Rhizobium and Bradyrhizobium, the microsymbionts of the

legume N2 fixing symbiosis, are well known microbial enhancement products

adopted across conventional, broad scale agriculture (Kahindi et al., 1997).

However, there are also a range of free living N2 fixing bacteria (including

Azotobacter, Azospirillum, Nitrobacter) that occur in the soil (and potentially also

stems and leaves) and that can actively fix N2, albeit typically at lower rates per

unit area than symbiotic fixation (Döbereiner & Pedrosa, 1987; Franche et al.,

2009; Kahindi et al., 1997) given access to a carbohydrate source and (generally)

micro aerophilic conditions. Such conditions may prevail in the rhizosphere.

Janzen et al. (1995) reported that pure isolates of N2 fixing Azospirillum,

amended with compost extract in a soil microcosm fixed more N2 than isolates

without compost extract. This suggests that compost extract stimulates microbiota

involved in nutrient cycling.

40
Chapter 1: Literature review

While less productive than a legume in terms of N2 fixation, free living N2

fixers have been reported to be responsible for significant rates of N2 fixation

(e.g. 75 kg N ha-1 year-1) (Jones & Bangs, 1985; Vadakattu & Paterson, 2006).

Some microbial enhancement products contain free living N2 fixing

bacteria. For example, the compost extract inoculant ‗Living Soil‘ (from

SmartBugs, Lismore, NSW) carries a claim of containing diverse species of N2

fixing microorganisms (e.g. Azotobacter spp., Azospirillum brasilense and

Bacillus spp.). This area has not received scientific attention, insomuch as there

are few reports in the literature of the efficacy of compost extract application in

enhancing soil N2 fixation (Janzen et al., 1995).

8.6 Plant disease suppression

Certain soil microbes have the ability to suppress numerous plant diseases

(Adams, 1990). Weltzien (1991) reviewed the experimental evidence related to

the suppressive effects of a variety of water-based compost preparations. It was

concluded that watery fermented extracts of different composts can suppress,

both in vivo and in vitro, a wide range of plant-pathogenic fungi. Several leaf

diseases, such as grey mold, apple scab, powdery mildew, downy mildew have

been controlled by compost extracts, prepared from different organic sources, and

as effectively as the conventional fungicides (Cronin et al., 1996; Hoitink et al.,

1997; Ketterer et al., 1992; Litterick et al., 2004; McQuilken et al., 1994;

Weltzein & Ketterer, 1986). Drench application of compost extracts has also been

reported to control soil-borne diseases (Litterick et al., 2004; Scheuerell &

Mahaffee, 2004).

41
Chapter 1: Literature review

The beneficial microorganisms in compost extract may act upon

pathogenic ones through: (i) direct competition (for the pathogen‘s resources,

such as nutrients, oxygen, space and water), (ii) antibiosis (inhibition or

destruction of the pathogen by metabolic products of the antagonist), or (iii)

parasitism (of the pathogen by the antagonist) (Brinton, 1995; Scheuerell and

Mahaffee, 2002). Management of the soil microbial population, including

addition of soil amendments, is a promising strategy for developing natural

inhibition of soil-borne diseases, thus improving crop production (Mazzola,

2004).

For example, incidence of Verticillium wilt and common scab of potato

has been reported to be reduced by the application of organic amendments

(Bailey & Lazarovits, 2003), and enhancement of growth and yield of okra has

been demonstrated for crops treated with compost extract enriched with

Trichoderma spp., on the premise that Trichoderma spp. are strongly antagonist

to other fungi (Siddiqui et al., 2008b). Similarly, Scheuerell and Mahaffee (2004)

reported that compost extract could be used as a drench in a soil-less container

medium to control damping-off in cucumber (cv. Marketmore 76) seedlings

caused by Pythium ultimum. An aerated compost extract amended with kelp and

humic acid was reported to be more effective than a non-aerated compost extract.

Also, Hibar et al. (2006) reported that in vitro application of compost extract

inhibited mycelium growth of Fusarium oxysporum f. sp. radicis-lycopersici and

also showed improvement in root and vegetative growth of an infected tomato

crop. Non-aerated compost extract has also been recommended in the context of

disease suppression (Weltzien, 1990). Minimum effective fermentation periods of

as short as one day (Urban & Trankner, 1993) have been reported for this

42
Chapter 1: Literature review

application, although Scheuerell and Mahaffee (2006) reported that 14 day old

non-aerated compost extract was more effective than 7 day old non-aerated

compost extract in suppressing Botrytis cinerea.

A summary of reports available in the scientific literature on disease

suppression by compost extract application is presented in Table 1.4.

43
Chapter 1: Literature review

Table 1. 4. Summary of reports on proven disease suppression following application of compost extracts (CEs) produced from known input materials
Author Pathogen Crop Dilution ratio / CEs Rate of Compost Methodology Comments
(Disease) application (Country)

Larkin (2008) Rhizoctonia solani Potato - 100 mL/kg Vermicompost containing USA Applied to soil
(Stem canker) composted horse manure,
coffee grounds, pepper, straw, clay
containing nutrient additives,
bluegreen algae, kelp, sugar and
yeast
Al-Mughrabi et al. Streptomyces scabies Potato - 140 L/ha - Field plot Applied to soil
(2008) (Common scab) (ACE) Canada

Haggag and Saber Alternaria solani Tomato 1:5 v/v 10 mL/pot Rice ash, bean straw and vegetative Greenhouse Applied to foliage
M.S.M.(2007) (Early blight) Onion (ACE and NCE) food waste, chicken manure Egypt
Alternaria porri
(Purple blight)

Welke (2000) Botrytis cinerea Strawberries 1:8 and 1:4 v/v 1.3 L/m2 Cattle compost and chicken Field plot Applied to soil
(Fruit rot) Broccoli (ACE and NCE) manure compost Canada
Rhizoctonia solani
(head rot)

El-Masry et al. Pythium debaryanum - 1:2 w/v - Leafy fruit compost, crop compost, In situ, in vitro Applied to Petri plates
(2002)a (Dieback) (CE) garden compost Egypt and culture broth
Fusarium oxysporum
f.sp. lycopersici
(Fusarium wilt)
Sclerotium bataticola
(Charcoal rot)
Utkhede and Koch Clavibacter michiganensis Tomato 1:21 L w/v - Kelp In vitro and greenhouse Applied to foliage on
(2004) (Bacterial canker) (CE) Canada pinched cotyledons

Scheuerell and Pythium ultimum Cucumber 1:1 and 1:4 v/v - Yard trimming, vermicompost extract Hydroponics Applied to soilless container
Mahaffee (2004)b (Damping off) (ACE) compost USA medium drench

Olanya and Larkin Phytophthora infestans Potato - - - Laboratory and Applied to Petri plates
(2006)c (Late blight) (ACE) growth chamber and potato foliage
USA
Dianez et al. 9 fungal pathogens 1:3 w/v 5, 10 and Grape marc (grapevine waste) compost In vitro Applied to Petri plates
-
(2006)d ( ACE) 15 % v/v Spain

a
in situ and in vitro experiments are done applying various concentrations.
b
compost extract was diluted by mixing compost in the ratio of 1:1, 1:4 and 1:9 with tap water and applied to the seeded pots.
c
effect of compost extract was found to be very feeble here (only 0-15% disease suppression) against Phytopthora infestans.
d
The pathogens used were: Rhizoctonia solani, Fusarium oxysporum (4 strains), Verticillium dahliae, Pythium aphanidermatum, Phytopthora parasitica and Verticillium fungicola

44
Chapter 1: Literature review

Table 1.4. continued…..


Author Pathogen Crop Dilution ratio / CEs Rate of Compost Methodology Comments
(Disease) application (Country)
Kone et al. (2010) Alternaria solani Tomato 1:5 v/v 15% v/v Composts from sheep manure, In vitro, greehouse Applied to Petri plates
(Early blight) (NCE) spray until runoff chicken manure, bovine manure, Canada and tomato foliage
Botrytis cinerea shrimp powder or seaweed
(Gray mold)
Phytophthora infestans
(Late blight)
Oidium neolycopersici
(Powdery mildew)
Siddiqui et al. (2009) Choanephora cucurbitarum Okra 1:5 w/v 0.1 mL /Petri plate Rice straw, empty fruit bunch In vitro and in vivo Applied to Petri plates
(Wet rot) (ACE) 1:1 CE to of oil palm composts Malaysia and sterilised PDA
sterilised PDA

Zmora-Nahum et al. (2008) Sclerotium rolfsii - 1:2 w/w 4.5 mL/Petri plate Compost mix of municipal In vitro Applied to Petri plates
(Southern blight) (Centrifused and filtered) sewage sludge and green Israel
waste (1:1 v/v) compost

Joshi et al. (2009) Phaeoisariopsis griseola French bean 1:5 v/v 100%, 1000 L/ha Composts of poultry manure, Field experiment Applied to foliage of
(Angular leaf spot) (NCE) farmyard manure, vermicompost, India French bean
spent mushroom compost,
Lantana camara and Urtica spp.
PDA - Potato dextrose agar.

45
Chapter 1: Literature review

Scheuerell (2003) reported that there was no difference between use of

aerated or non- aerated compost extract in control of powdery mildew on field-

grown roses, in control of grey mold on greenhouse grown geraniums, or in

control of damping-off of basil seedlings indicating the lack of difference in

microbial density and diversity between those compost extracts.

Larkin (2008) evaluated aerated compost extract in the field and

greenhouse alone and in combination with different rotations for its efficacy to

enhance the soil microbial community and to suppress soil-borne diseases of

potato. In the barley/ryegrass rotation, he found that soil application of aerated

compost extract combined with a beneficial microorganism mix of Trichoderma

griseoviridis, Bacillus spp., Trichoderma harzianum and organic nutrients (humic

acids, kelp and yeast) reduced diseases, such as stem canker, black scurf and

common scab separately by 18-33% on tubers and increased yield by 20-23% but

not in other rotations. These results indicated that soil amendments generally add

or enhance microbiota in the soil and the strong effect of crop characteristics to

shape the microbial community is possibly the reason for developing disease

suppressive soils.

Various parameters of the compost extract production process were tested

for suppression of grey mold (Botrytis cinerea) on geranium (Scheuerell &

Mahaffee, 2006) and it was shown that compost extract produced with continuous

aeration did not suppress grey mold disease whereas the non-aerated compost

extract did. Although not consistent, when the aerated compost extract with

continuous aeration was mixed with kelp extract, rock dust and humic acid, it

reduced the disease significantly. The reason for this was unclear because

46
Chapter 1: Literature review

suppression of grey mold was not matched by an increase in the number of

cultivable bacteria in the extract.

Organic compost extracts prepared from pig, horse and cow manure were

tested against fusarium wilt of sweet pepper (Fusarium oxysporum f. sp.

vasinfectum) and all the treatments were effective in suppressing this pathogen by

up to 88.5% (Ma et al., 2001b). Compost water extract (non-aerated) prepared

from cattle manure successfully reduced tomato leaf grey mold (B. cinerea) as

opposed to water control in a greenhouse assay, however the best suppression

was achieved with the chemical fungicide, vinclozolin (Elad & Steinberg, 1994).

8.7 Acquired systemic resistance

A good example of acquired systemic resistance due to the non-

pathogenic colonisation has been given by a team of researchers at the University

of Delaware, U.S. (Bryant, 2008). They found that when the plant leaf is attacked

by a pathogen, the plant recognises the attack and sends a signal to the root

system. The root responds by secreting organic acids into the rhizosphere which

attract beneficial bacteria to help protect the plant by inducing systemic resistance

against the disease causing pathogens. They reported that the leaves of the plants

(the flowering plant, Arabidopsis thaliana) infected with a pathogenic bacterium

(Pseudomonas synngae) but without root inoculation of the beneficial microbe

(Bacillus subtilis) showed chlorosis and other disease symptoms, however the

same disease-infected plants but with root inoculation (B. subtilis) remained

healthy.

Several greenhouse and field studies have shown plant induced systemic

resistance against a variety of pathogens when cucumber seeds were treated with

47
Chapter 1: Literature review

plant growth promoting rhizobacteria (Liu et al., 1995; Raupach & Kloepper,

1997; Raupach et al., 1996). Wei et al. (1991) also reported induced resistance

response against Colletotrichum orbiculare in the leaves of cucumber when

rhizosphere bacteria were applied to their seeds. Such growth promoting

rhizobacteria have demonstrated a great potential in biological control of several

diseases (Kloepper et al., 1996). Similarly, Pseudomonas spp. was reported to

induce systemic resistance in carnation to fungal wilt caused by F. oxysporum

and accumulated increased concentration of phytoalexins in the carnation stem

(van Peer et al., 1991).

Biocontrol of plant diseases is a complex system and its understanding

and exploitation requires a deep insight of several sciences, including microbial

ecology, system analysis, epidemiology, biochemistry and molecular

biotechnology (Elad et al., 1996).

8.8 Bioremediation

Bioremediation is the term given based on the activity of introduced

living organisms to the restoration of an area contaminated with, typically, a

chemical pollutant. Various techniques have been developed in the past to

rehabilitate polluted soils (Alexander, 1994; Romantschuk et al., 2000; Skladany

& Metting, 1992). The chemical pollutant may be a metal (e.g. arsenic associated

with old cattle dips; or aluminium ion in acid soils), a hydrocarbon (e.g. a diesel

oil spill), or other classes of organic compounds (e.g. herbicide residues), or

chemical compounds (e.g. cyanide from ore processing and manufacturing

processes). A variety of microbial consortia that are capable of degrading

environmental pollutants (e.g. bacteria, fungi, grazing protozoa) can be involved

48
Chapter 1: Literature review

in the bioremediation process (Watanabe, 2001). The process of bioremediation

may involve a single organism, or may involve a succession of communities of

organisms. Organic wastes, such as petroleum can be effectively biodegraded

using indigenous organisms in the presence of suitable nutrients and an oxygen

supply (Allard & Neilson, 1997). Microbial degradation of cyanide has been

demonstrated and species known to act in cyanide degradation included

Pseudomonas fluorescens, Trichoderma koningii, Fusarium oxysporum, Bacillus

stearothermophilus, Staphylococcus seiuri, and so on (e.g. Dursun & Aksu, 1999;

Ezzi & Lynch, 2002; Gurbuz et al., 2009).

Bioremediation of chemical pollutants in soil is seen as a cost-effective

and environmentally acceptable approach (Buswell & Eriksson, 1994). The

effectiveness of a soil in bioremediation will be partly physio-chemical (retention

of the material within the soil, catalytic conversion to a non-toxic form) and

partly biological. Microbially enhanced compost extract (‗MECE‘ or ‗CE‘) may

increase relevant soil microbiota and may contribute chemical components

relevant to this issue.

While no peer-reviewed journal article was found on the topic of the

efficacy of compost extract in bioremediation, there is published work on the

remediation of polluted soil using compost (e.g. Diatloff et al., 1998; Dumestre et

al., 1999). The use of compost extract to remediate contaminated soil is a logical

extension.

A potential area for use of compost extract could be biostabilisation or

detoxification of the toxic trivalent aluminium ion. Aluminium toxicity arises

from the high solubilisation of aluminium with decreases in soil pH, a serious

factor causing yield reduction in many acidic soils (Wright, 1989). Previous

49
Chapter 1: Literature review

research has shown that humic and fulvic acids can complex aluminium, reducing

trivalent monomeric aluminium in nutrient solutions, with a resultant increase in

the root length of corn (Diatloff et al., 1998). Compost extracts should contain

humic and fulvic acids, and so should be able to biostabilise aluminium in the

rhizosphere.

9. Summary and hypotheses

Given this interest, the field claims for the benefit of compost extract, and

the need for a theoretical basis for such claims, it is timely for further scientific

investigation of this product and practices involved in its use. While not a

‗complete solution‘, it is likely that this technology may be a useful additional

tool to support ‗biological farming‘ and sustainable agriculture.

Farmer interest in, and uptake of, compost extract use is increasing. This

change is underpinned by interest in sustainable farming practices (building soil

‗health‘) and in lowering the cost of inputs (particularly driven by rising fertiliser

prices). There is a diverse array of potential benefits claimed for the use of

compost extract. While most of these benefits can be generally understood in

terms of mechanism of action in the context of current soil, plant and microbial

science, some claims are less explicable. As an example of the latter, how can

compost extract enhance microbiota generally while inhibiting some microbiota

specifically (i.e. plant pathogens)? Likewise, what is the mechanism behind the

claim that compost extract can cause the plant to develop systemic resistance (i.e.

that compost extract applied to plant roots can result in increased plant resistance

to leaf pathogens)?

50
Chapter 1: Literature review

While mechanisms for compost extract action can be hypothesised, there

is, unfortunately, a paucity of consideration of this topic in the scientific

literature. For example, Carpenter (2005) notes the need to explore for a ―base

theory‖ about compost extract, in terms of type of microbial community,

classification of beneficial microbes, different conditions for incubation and

subsequent effects on microbes. Research to date has largely focused on the areas

of plant nutrition and disease suppression, but this has generally been at an

empirical field level with little consideration of the underlying mechanism of

action.

There is a very limited published work on the characterisation of compost

extract. This is unfortunate, given that compost extract may be a highly variable

product. This variation may be a major factor in reports of inconsistency in plant

response to compost extract application, with reports ranging from highly

beneficial to complete failure, especially from a disease suppression point of view

(Bess, 2000; Larkin, 2008; Siddiqui et al., 2008a).

Based on the literature review, the following hypotheses could be

addressed in understanding the possible benefits of compost extract to plants:

(a) Application of compost extract to the soil can directly improve plant

nutrition through nutrients derived from the inoculant compost and other

amendments added to the compost extract.

(b) Application of compost extract to the soil and leaf surfaces at low rates (e.g.

20-50 L ha-1) can result in a persistent change in soil and phylloplane

microbial populations.

(c) Application of compost extract to the soil can indirectly improve plant

nutrition through mineralisation of the soil organic fraction. This benefit may

51
Chapter 1: Literature review

especially apply to P nutrition, through P-solubilising bacteria and increased

mycorrhizal density.

(d) Above-ground application of compost extract can accelerate the degradation

rate of leaf litter and standing stubble.

(e) Application of compost extract to the soil can increase N2 fixation by free

living soil bacteria.

(f) Application of compost extract to the soil and to above-ground biomass can

result in suppression of root and foliar plant diseases, respectively, through

competition, antibiosis or parasitism, by compost extract microbiota, of the

pathogen.

(g) Application of compost extract to one part of the plant can result in induction

of systemic resistance throughout the plant.

(h) Application of compost extract to the soil can result in bioremediation of

toxic organic and inorganic species present in the soil. For example, compost

extract may allow increased root growth in the presence of Al+++ ions via

chelation of these ions by humic acids or sequestration by micobiota.

(i) Application of compost extract to the soil can result in improved soil C

content through addition of organic matter via partial decomposition of plant

residues.

(j) Application of compost extract to the soil can result in improved soil

structure through direct addition of humates to the soil and through addition

of organic matter via partial decomposition of plant residues.

(k) Application of compost extract to the soil can result in improved water

holding capacity through improvement in soil structure and through addition

of organic matter via partial decomposition of plant residues.

52
Chapter 1: Literature review

(l) Long-term application of compost extracts to the soil can alter soil

microbiota at the field level and influence soil fertility and the possibility of

building a disease suppressive soil.

10. Conclusion

Given the growing interest in using compost extract, as a sustainable

management practice, with the consideration of its huge potential benefits in crop

production including commercial cultivation, and the lack of scientific research

on its possible roles in agriculture, it is timely to consider and evaluate the

potential effects of compost extracts on crop production. Attention should also be

given to the employment of various practices including microbiological and

molecular methods to be used for the monitoring of microbial populations in the

trials. The findings of this research may provide an insight to all users/producers

of microbially enhanced compost extracts including small and large scale farmers,

and researchers.

To achieve this, a number of experiments were conducted on production

and characterisation of compost and of compost extracts. Based on the literature

review, the potential contribution of compost extract to soil ‗health‘ may be

through, although not limited to: (a) soil mineral solubilisation, (b) organic matter

turnover (mineralisation), (c) aluminium chelation, and (d) disease suppression.

These claims were experimentally tested and explanations for the mode of each

effect have been sought.

53
Chapter 2: Compost based extract - characterisation

Chapter 2

Microbial enhancement of compost extracts based

on cattle rumen content compost – characterisation

of a system

A version of this chapter has been published in Bioresource Technology by Karuna Shrestha,
Pramod Shrestha, Kerry B. Walsh, Keith M. Harrower and David J. Midmore

Abstract

Microbially enhanced compost extracts (‗compost tea‘) are being used in

commercial agriculture as a source of nutrients and for their perceived benefit to

soil microbiology, including plant disease suppression. Rumen content material is

a waste of cattle abattoirs, which can be value-added by conversion to compost

and ‗compost tea‘. A system for compost extraction and microbial enhancement

was characterised. Molasses amendment increased bacterial count 10-fold, while

amendment based on molasses and ‗fish and kelp hydrolysate‘ increased fungal

count 10-fold. Compost extract incubated at 1:10 (w/v) dilution showed the

highest microbial load, activity and humic/fulvic acid content compared to other

dilutions. Aeration increased the extraction efficiency of soluble metabolites, and

microbial growth rate, as did extraction of compost without the use of a

54
Chapter 2: Compost based extract - characterisation

constraining bag. A protocol of 1:10 dilution and aerated incubation with kelp

and molasses amendments is recommended to optimise microbial load and

fungal-to-bacterial ratio for this inoculum source.

1. Introduction

Composting can be viewed as an organic waste reduction process (Eiland

et al., 2001), but its greater value lies in the use of compost as a soil conditioner

(Leu, 2006). For large scale composting, point sources of organic material of

uniform quality are required. The two abattoirs in Rockhampton (Qld, Australia)

slaughter in total approximately 1700 head a day, producing 140 cubic meters of

rumen content per day (R. Lang, 2007; pers. comm.). Rumen content material is

composed of partly digested plant material. This waste has been historically

disposed across neighbouring grazing properties, but in recent years it has been

composted and sold. A windrow composting process is undertaken over 9

months, with windrows turned mechanically at monthly intervals in order to

provide aeration and improve homogeneity.

A water-based preparation of compost, incubated with various microbial

food sources and in some cases inoculated with specific microbiota, is defined as

a ‗microbially enhanced compost extract‘ or ‗compost extract‘ in short (and

popularly termed as ‗compost tea‘). Compost extracts are gaining popularity,

particularly, amongst those who are seeking substitutes to agrochemicals (Bess,

2000). ‗Mainstream‘ growers are also looking to reduce petrochemical-derived

products such as fertiliser, as prices soar.

Compost extract contains a high population of microbiota, e.g.

Rhizobacteria, Trichoderma, and Pseudomonas spp., which may enhance growth

55
Chapter 2: Compost based extract - characterisation

and yield of crops (Sylvia, 2004). These microbiota produce plant growth

hormones and chemical compounds (e.g. siderophores, tannins, phenols) which

are antagonistic to various soil pathogens (Antonio et al., 2008). Other microbiota

may benefit plants through mechanisms such as nitrogen fixation and phosphate

solubilisation (Dubeikovsky et al., 1993). The use of compost extract is also

claimed to increase soil C levels, improve soil structure, nutrient cycling and

water holding capacity, and suppress plant diseases (Ha et al., 2008). However, to

achieve these benefits, several variables have to be considered to produce

compost extracts of desired quality. These include microbial food sources,

compost to water ratio, levels of aeration, compost quality, compost age, duration

of incubation, and the quality of water used. There is also a need for consistent

compost quality, which depends on consistency of inputs and methods used to

produce compost.

Given that the claims for the benefits of compost extracts are varied (from

disease suppression, and improved soil nutrient cycling to a simple direct plant

nutrition effect), the definition of compost extract ‗quality‘ is open ended. For

this study, we assume that the process goal is to maximise nutrient extraction

from compost, and to maximise microbial growth, particularly fungal growth.

Sugar or molasses, kelp extract, fish emulsion and rock dust are

commonly used as cost effective substrates usually available in bulk (Ingham,

2005). Addition of such microbial food sources to the compost extract will

increase microbial population growth during the incubation period (Naidu et al.,

2010). The additives used will affect the C-N balance, the form of C

(carbohydrate), and the form of the N source in growth media and this in turn will

affect the species composition (e.g. fungal: bacterial ratio) of the culture. Pant et

56
Chapter 2: Compost based extract - characterisation

al. (2009) indicates that further work is required to relate the impact of such

additives on extract quality (in terms of minerals and microbiota).

Compost to water ratio is an important factor influencing the nutritional

and microbial status of the final product (Weltzien, 1990). Researchers have

employed a wide range of dilution ratios ranging from as low as 1:1 (Zhang et al.,

1998) to as high as 1:50 (Weltzien, 1990). Although dilution ratios of 1:3 to 1:10

are common (Scheuerell and Mahaffee, 2004), ratios of up to 1:1000 are currently

practiced in agriculture and horticulture (D. Daniels, 2008 pers. comm.). The

dilution ratios will obviously impact the nutrient concentration and microbial load

of the final product. There is also increased risk of biological contamination in

heavily diluted extracts.

Compost extract can be produced under aerobic or anaerobic conditions,

with obvious consequences to the microbial community supported. Past research

has mostly focussed on the production and application of non-aerated compost

extracts (Elad & Steinberg, 1994), but increasing attention is being given to

aerobically produced compost extracts, particularly for the control of a variety of

plant diseases (e.g. Siddiqui et al., 2009).

Filtering of compost extracts is required to prevent clogging of emitters or

spray nozzles (Weltzien, 1991). To avoid separate filtration steps, many users

incubate compost contained in a permeable bag (e.g. made up of muslin)

(Ingham, 2005). This practice can be expected to reduce diffusion between the

compost and the bulk solution, and so reduce the rate of microbial and nutrient

extraction from the compost. Whether this practice impacts on species

composition of the extract is not known.

57
Chapter 2: Compost based extract - characterisation

The production and use of compost extracts for agronomic and

horticultural purposes appears to be gaining in popularity. For example, in the

central Queensland, compost extracts are used in horticultural, broadacre grain

and grazing enterprises, either produced on-farm or supplied by contractors. A

few scientific studies have been undertaken that have evaluated compost extracts

in terms of nutrients and microbiota (Naidu et al., 2010; Pant et al., 2009;

Scheuerell, 2004; Scheuerell & Mahaffee, 2006). Further studies are essential to

understand the impact of specific incubation conditions on the microbial

outcomes. In this study, the impacts of four major production variables, namely,

additives, dilution rate, level of aeration and use of bags during the incubation of

compost extract are investigated.

2. Materials and methods

2.1. Experiments and treatments

The experiments were carried out indoors at ambient temperatures

between a range of 22 to 32˚C. Compost extractions were undertaken with

variation in (i) additives (experiment 1), (ii) compost: water ratio (experiment 2)

and (iii) aeration and use of a bag to contain the compost (experiment 3). Each

experiment consisted of four treatments with three replications, utilising 12

buckets of 60 L size. A completely randomised design was used for subsequent

experiments. The default extract condition consisted of 3 kg (f wt) compost

contained in a cotton (28 g m-2) bag, placed in 30 L of town water (i.e., a 1:10

ratio) and continuously aerated at the rate of 36 L min-1 through air stones located

at the base of the bucket, as per Shrestha et al. (2011a). The water was aerated

58
Chapter 2: Compost based extract - characterisation

vigorously for 30 minutes to remove residual chlorine prior to initiation of

treatment.

The first experiment involved aerated incubation of compost extracts with

two different additives, namely, ‗fish and kelp hydrolysate‘ (kelp) from Searle

Pty Ltd, Australia, molasses (molasses), ‗fish and kelp hydrolysate‘ + molasses

combination (kelp + molasses) and a no additive treatment (control). ‗Fish and

kelp hydrolysate‘ and/or molasses were added at the rate of 1% and/or 0.5%

(v/v), respectively, except in the control. ‗Fish & kelp hydrolysate‘ consists of

78% fish hydrolysate, 20% kelp, 2% stabiliser and 4% fish oil on w/w basis and

contains 10.3% N, 2.5% P, 2.5% K, 7% Ca, 0.2% Mg, 0.5% S, 690 mg kg-1 Fe, 5

mg kg-1 Cu, 14 mg kg-1 Mn, 32 mg kg-1 Zn, 14 mg kg-1 B, and 1 mg kg-1 Mo in

organic form. Molasses is a by-product of cane processing (simple sugars).

The second experiment consisted of compost extracts incubated in

different dilution ratios (non-diluted, 1:10, 1:100, and 1:1000 compost:water on a

w/v basis) with all incubations involving kelp and molasses additions under

aerated conditions.

The third experiment consisted of aerated (ACE) and non-aerated

compost extract (NCE) treatments incubated without or with (ACEb and NCEb)

the use of cotton bags to retain the compost during incubation. Here also, kelp

and molasses were added to each treatment immediately after addition of

compost.

2.2. Source compost

Compost based on cattle rumen content was obtained from a commercial

composting operation (Broadmeadows Pty Ltd). In the Rockhampton area, most

59
Chapter 2: Compost based extract - characterisation

animals are fed by grazing (rather than feedlot). All three experiments utilised

freshly collected rumen compost (nine month old) as a basic source material to

produce compost extracts. As there can be considerable batch-to-batch

inconsistency in compost composition (Campbell, 2006), a single batch of

compost was used to produce the extract in each experiment.

2.3. Physico-chemical analyses

In all treatments, the extract pH, electrical conductivity (EC), temperature

and dissolved oxygen, as well as ambient temperature were determined at 3 h

intervals over a 24 h period and a final reading was taken at 48 h. The pH and EC

were measured using a TPS labCHEM Cond/pH meter while dissolved oxygen

and temperature of the extracts were measured using a TPS WP 82 Dissolved

Oxygen Meter (EnviroEquip, Australia).

Samples of compost extracts were collected after 24 h of incubation and

concentrations of inorganic soluble ions (NO3−-N, NH4+-N, PO4−-P and K+-K)

were determined in triplicate using colorimetric test strips (RQflex plus).

Humic and fulvic acids were determined from samples collected at 24 h

using a UV-Vis spectrophotometer following a slightly modified method by

Yamada et al. (1998). Briefly, 2 mL of centrifuged and filtered (0.45 µm

membrane filter) sample of compost extract was added to 8 mL of reverse

osmosis (RO) water followed by addition of 10 M hydrochloric acid (500 L). The

precipitated humic acid was separated from the supernatant solution by

centrifugation at 7000 rpm for 72 minutes at 4˚C. The supernatant solution was

neutralized by adding 10 M sodium hydroxide and the absorbance was measured

at 350 nm using UV spectrophotometer to index fulvic acid level. The

60
Chapter 2: Compost based extract - characterisation

precipitated humic acid was dissolved with 1M sodium hydroxide followed by

neutralization with 1 M hydrochloric acid. The absorbance was then measured at

350 nm to index humic acid level. The humic and fulvic acid concentrations of

the samples were then calculated by reference to humic (Sigma Aldrich,

Switzerland) and fulvic acid (International Humic Substances Society, USA)

standards, respectively.

2.4. Microbial analyses

Compost extract samples were collected in triplicate after 24 h of

incubation and cultivable bacterial and fungal populations of serially diluted

samples were estimated by the pour plate technique using plate count agar for

bacteria and half strength potato dextrose agar with chloramphenical (100 µg mL-
1
) for fungi (Harrower, 2006). Bacterial plates were incubated at 37˚C for 1-2

days while fungal plates were incubated for 3-5 days at 25˚C.

Total microbial activity of the samples collected after 12, 24 and 48 h of

incubation were determined using the fluorescein diacetate (FDA) hydrolysis

method given by Adam and Duncan (2001).

2.5 Statistical analysis

Statistical analysis of non-transformed data was undertaken by two-way

ANOVA using the statistical package (XLSTAT, 2010) using GLM model.

Differences between means were separated by Tukey‘s test (p< 0.05). Data in

figures is presented as a mean with associated standard error, with indication of

significant differences between means for a given time of treatment.

61
Chapter 2: Compost based extract - characterisation

3. Results and discussion

3.1. General trends

The three experiments consisted of a common treatment with 1:10 w/v

compost contained within a muslin bag in an aerated solution containing

molasses and kelp as additives. Similar ammonium, phosphate and potassium

levels were recorded for this treatment, indicative of a consistent compost source,

although the first experiment produced an extract relatively richer in humic acid,

and the third experiment produced an extract lower in microbial population,

although with a similar bacterial to fungal ratio (Tables 1-3). While these

differences could be due to changes in growing conditions, it is most likely they

represent changes in compost source, reinforcing the need to use compost

produced under consistent conditions.

Adegunloye et al. (2007) reported that most of the microorganisms in

compost best survived under neutral pH, and presumably this is also true for

compost extract. On this basis, a pH of 7 is ideal for compost extracts. Increasing

the EC of an extract is indicative of increased extraction of anions and cations,

which should increase the benefit of the extract as a plant fertiliser; however, an

EC above 5 dS m-1 is likely to decrease microbial survival (Okur, 2002) and is

thus an upper quality limit for compost extracts. In general, the pH of the

compost extract increased during incubation to around 6, and EC increased to 4

dS m-1. The increase in pH is likely due to an overall uptake of more anions than

cations by the microbiota (as commonly documented in agricultural systems, e.g.

Yan et al., 1996) while the increase in EC presumably represents leaching and

breakdown of the organic matter of the compost.

62
Chapter 2: Compost based extract - characterisation

Solution temperature and oxygen levels are indices of microbial activity

(e.g. as reported by Garcia et al., 1991 for compost). Tiquia et al. (1996)

correlated temperature with dehydrogenase activity, ATP content and oxygen

consumption rate during composting of spent litter from pig pens at differing

moisture levels and reported that these parameters are indicative of microbial

activity. During incubation, temperature increased above ambient and solution

oxygen levels declined dramatically around 9 h after start of incubation. The

latter decrease indicates that the aeration method failed to keep pace with the

oxygen demand of the culture, despite these liquid cultures being aerated

continuously at the rate of 36 L air min-1 per vessel (30 L). Oxygen depletion to

levels below 6 ppm for extended periods may alter microbial composition of the

culture (Ingham, 2005). This depletion could be addressed by either improving

the aeration system or by decreasing microbial activity (e.g. less microbial food

per bucket, or gradual addition of microbial food during incubation).

3.2. Experiment 1: Comparison of compost extracts with different

additives (control, kelp, molasses and kelp+molasses)

The pH and EC of a 1:5 dilution (with distilled water) of molasses and

kelp was 5.5 and 4.3, and 2.3 and 1.0 dS m-1, respectively. The additives lowered

the pH of compost extracts, and increased extract EC. Significantly higher pH

values (Fig. 2.1A) were recorded in the control (no additives), while the kelp +

molasses treatment showed the highest EC, followed by molasses treatment (Fig.

2.1B). Naidu et al. (2010) have reported that fortification with additives such as

kelp, peptone, humic acid and yeast extracts increased the pH, while brown sugar

and corn meal decreased the pH of compost extracts. They found that EC values

63
Chapter 2: Compost based extract - characterisation

of compost extracts were also increased with all the additives tested compared to

the no-additive control. Scheuerell and Mahaffee (2004) also found an increase in

pH values from 7.4 in the non-additive control to between 7.9 and 8.6 in compost

extracts produced with addition of a bacterial nutrient solution obtained from Soil

Soup Inc. WA, USA, and a fungal medium based on soluble seaweed powder,

liquid humic acids and rock dust, adapted from Ingham and Alms (1999). In

summary, in the non-buffered solution of compost extract, pH of an additive is a

major determinant of the pH of a final extract.

64
Chapter 2: Compost based extract - characterisation

Figure 2. 1. Comparison of pH, electrical conductivity (dS m-1), fluctuations in temperatures (˚C)
and dissolved oxygen (% of saturation) over time (readings taken at every 3 h intervals until 24 h
and final readings taken at 48 h) in rumen compost extracts with different additives. Control -
compost extract with no additives, Kelp - compost extract with ‗fish and kelp hydrolysate‘,
Molasses - compost extract with molasses, Kelp+Molasses - compost extract with ‗fish and kelp
hydrolysate‘ and molasses. Values are mean ± SE (n = 3).

65
Chapter 2: Compost based extract - characterisation

The difference between EC values at the start of incubation must

represent difference in EC of additives. The rise in EC with incubation time must

represent extraction of ions from the compost. The rapid rise within 3 h, followed

by a plateau, indicates that mixing conditions were adequate to ensure solution

diffusion through the compost bag.

The inorganic phosphate and potassium concentrations did not differ

among the treatments; however, the concentrations of ammonium in the molasses

and kelp + molasses treatments were lower than those of the control (Table 2.1).

In contrast, Pant et al. (2009) found higher levels of nitrate, ammonium and

potassium after adding kelp extract and humic acid to their original liquid

preparation. Similarly, Naidu et al. (2010) reported significantly higher

concentrations of nitrogen, phosphorus and potassium along with iron, calcium

and manganese in microbially enriched compost extracts compared to the no

additive control. Specifically, the authors used yeast extract and humic acid as the

source of nutrients. In our study, the increased microbial activity associated with

the addition of molasses and kelp presumably resulted in mineralisation of some

organic forms into mineral forms in the solution. The lower ammonium level in

extracts amended with additives could be due to increased volatilization (for pH

below 6) or to uptake by the increased microbiota present in these treatments.

Although no difference in humic acid was noted, fulvic acid level was

significantly higher in the kelp treatment compared to the control treatment (0.8

vs. 1.23 µg g-1 d wt), which suggests that ‗fish and kelp hydrolysate‘ was a

significant source of fulvic acid (Table 2.1).

66
Chapter 2: Compost based extract - characterisation

Table 2. 1. Comparison of physico-chemical and microbial characteristics of compost extracts


incubated with different additives and sampled after 24 h of incubation.

Parameter Control Kelp Molasses Kelp+Molasses


+ -1
NH4 -N (µg g d wt) 1.32 ± 0.04b 1.19 ± 0.03ab 1.06 ± 0.07a 1.09 ± 0.04a
- -1
PO43 -P (µg g d wt) 32.83 ± 2.27a 30.44 ± 2.14a 29.35 ± 0.75a 32.39 ± 0.43a
+ -1
K -K (µg g d wt)-1 0.55 ± 0.00a 0.59 ± 0.02a 0.61 ± 0.03a 0.69 ± 0.06a
Humic acid (µg g d wt) 2.33 ± 0.17a 2.08 ± 0.46a 2.17 ± 0.03a 1.44 ± 0.30a
-1
Fulvic acid (µg g d wt) 0.82 ± 0.08a 1.23 ± 0.07b 0.96 ± 0.07ab 0.98 ± 0.07ab
Bacterial population (cfu Log10) @ 24h 10.76 ± 0.07b 7.78 ± 0.02a 11.49 ± 0.05c 10.64 ± 0.07b
Fungal population (cfu Log10) @ 24h 5.88 ± 0.02a 6.62 ± 0.02c 6.14 ± 0.05b 6.78 ± 0.01d
d wt Dry weight. Control - compost extract with no additives, Kelp - compost extract with 'fish and kelp hydrolysate',
Molasses - compost extract with molasses, Kelp+Molasses - compost extract with 'fish and kelp hydrolysate' and molasses.
Within rows, means (n = 3) with the same letter are not significantly different according to Tukey's test (p< 0.05).

The molasses treatment showed the highest temperature followed by kelp

+ molasses treatment (Fig. 2.1C), while these two treatments also demonstrated

the quickest decrease in the level of dissolved oxygen (Fig. 2.1D), presumably

reflecting microbial activity. The level of dissolved oxygen in each treatment

(Fig. 2.1D) was consistent with the activity of microorganisms at 48 h of

incubation (as assessed using FDA, Fig. 2.2) and plate counts of culturable

microorganisms (Table 2.1).

Figure 2. 2. Total microbial activity over time (readings taken at 12, 24 and 48 h) as estimated by
the FDA hydrolysis method in compost extracts with different additives. Control - compost
extract with no additives, Kelp - compost extract with ‗fish and kelp hydrolysate‘, Molasses -
compost extract with molasses, Kelp+Molasses - compost extract with ‗fish and kelp hydrolysate‘
and molasses. Within a time, treatment means with the same letter do not differ significantly and
values are mean ± SE (n = 3).

67
Chapter 2: Compost based extract - characterisation

Bacterial and fungal populations as indicated by the plate counts were in

the order molasses > kelp + molasses = control > kelp and kelp + molasses > kelp

> molasses > control, respectively. In agreement with the bacterial plate counts,

the lowest microbial activity was observed in the kelp treatment as also

demonstrated by FDA hydrolysis (Fig. 2.2), although the ranking of other

treatments (kelp + molasses = molasses = control > kelp) opposed to the plate

counts (Table 2.1). The difference between the order of results of the fungal plate

count method and the FDA hydrolysis method agrees with the finding by Garcia-

Gomez et al. (2003), and presumably reflects the difference between a measure of

bacterial activity (FDA method), as opposed to a measure of number of

propagules (plate counts).

Different microorganisms have specific substrate requirements, therefore,

it is expected that the types of organic compounds added to a culture will

influence microbial diversity. The highest bacterial population was achieved with

molasses amendment, while the highest fungal to bacterial ratio was achieved

with both kelp and molasses amendment (Table 2.1). The reason for this could be

due to the nature of the C substrate in each of these additives (molasses should be

simple sugar, while kelp should have included cellulose and lignified plant

tissues). Of the additives tested, the use of both molasses and kelp is

recommended.

3.3. Experiment 2: Comparison of compost extracts with different

dilution levels

Increasing the ratio of compost to water resulted in significantly higher

values of pH and EC. EC plateaued after 3-6 h of incubation even in the 1:10

68
Chapter 2: Compost based extract - characterisation

ratio treatment, indicating that a reasonable diffusion of solutes was occurring

through the compost into the bulk solution, despite its containment in a bag.

Compost extract produced using a ratio of 1:10 was also higher in concentrations

of ammonium, humic acid and fulvic acid compared to those of other extracts

(Table 2.2).

Table 2. 2. Comparison of physico-chemical and microbial characteristics of compost extracts


incubated with different dilutions and sampled at 24 h of incubation.

Parameter Non-diluted 1:10 1:100 1:1000


+ -1 0.18 ± 0.07a 1.29 ± 0.10b 0.28 ± 0.03a
NH4 -N (µg g d wt) 0.16 ± 0.04a
- -1 37.61 ± 3.20a 33.05 ± 1.52a 39.35 ± 1.15a 40.00 ± 2.27a
PO43 -P (µg g d wt)
+ -1
K -K (µg g d wt) 0.53 ± 0.03a 0.73 ± 0.02b 0.69 ± 0.06ab 0.63 ± 0.02ab
-1
Humic acid (µg g d wt) 0.12 ± 0.02a 0.60 ± 0.20b 0.08 ± 0.01a 0.09 ± 0.01a
-1
Fulvic acid (µg g d wt) 0.71 ± 0.03a 1.80 ± 0.10b 0.78 ± 0.07a 0.62 ± 0.07a
Bacterial population (cfu Log10) @ 24h 9.24 ± 0.04a 11.79 ± 0.01d 10.96 ± 0.04c 9.39 ± 0.03b
Fungal population (cfu Log10) @ 24h 7.84 ± 0.04a 8.40 ± 0.07b 8.28 ± 0.05b 8.21 ± 0.01b
d wt Dry weight. Non-diluted - compost extract with no compost inoculum but with 'fish and kelp hydrolysate'
and molasses, 1:10 - compost extract at the ratio of 1 part compost to 10 parts water (w/v), 1:100 - compost
extract at the ratio of 1 part compost to 100 parts water (w/v), 1:1000 - compost extract at the ratio of 1 part
compost to 1000 parts water (w/v). Within rows, means (n = 3) with the same letter are not significantly
different according to Tukey's test (p< 0.05).

Solution temperatures of the four compost dilution treatments did not

differ initially; however, after 12-15 h of incubation, temperatures of three levels

of dilution were significantly higher than that of the control treatment (Fig. 2.3C).

Dissolved oxygen decreased rapidly (to approx less than 10%) between 6-12 h

after culture initiation in the 1:10 treatment, and 9-15 h in 1:100 and over 15-21 h

in 1:1000, indicative of the level of microbial activity during incubation (Fig.

2.3D). Indeed, the compost acted as a bacterial inoculum, with higher bacterial

plate counts related to increasing compost ratios (Table 2.2). The compost was

also a source of fungal inoculum; however, the culturable fungal load in the 48 h

incubated extracts was not proportional to the compost:water ratio (Table 2.2).

69
Chapter 2: Compost based extract - characterisation

A 9
8
7
6
5

pH
4
3 Non-diluted
1:10
2 1:100
1 1:1000

0
0 3 6 9 12 15 18 21 24 48

5.0
4.5
4.0
3.5
EC (dS m-1)

3.0
2.5
2.0
1.5 Non-diluted
1:10
1.0 1:100
0.5 1:1000

0.0
0 3 6 9 12 15 18 21 24 48

C 30
29
28
Temperature (°C)

27
26
25
Non-diluted
24 1:10
1:100
23 1:1000

22
0 3 6 9 12 15 18 21 24 48

120
D
100
Dissolved oxygen (%)

80
Non-diluted
60 1:10
1:100
1:1000
40

20

0
0 3 6 9 12 15 18 21 24 48
Time (h)

Figure 2. 3. Time course of solution pH, electrical conductivity (dS m-1), temperatures (˚C) and
dissolved oxygen (% of saturation); readings taken at every 3 h intervals until 24 h and final
readings taken at 48 h in rumen compost extracts with different dilutions. Non-diluted - compost
extract with no compost inoculum but with ‗fish and kelp hydrolysate‘ and molasses, 1:10 -
compost extract at the ratio of 1 part compost to 10 parts water (w/v), 1:100 - compost extract at
the ratio of 1 part compost to 100 parts water (w/v), 1:1000 - compost extract at the ratio of 1 part
compost to 1000 parts water (w/v). Values are mean ± SE (n = 3).

70
Chapter 2: Compost based extract - characterisation

The higher level of oxygen demand in the 1:10 treatment (Fig. 2.3D) was

also in agreement with the highest relative total microbial activity as indicated by

the highest amount of fluorescein release (Fig. 2.4) and by the greater number of

culturable microorganisms (Table 2.2).

Figure 2. 4. Comparison of total microbial activity at different times (readings taken at 12, 24 and
48 h) following the FDA hydrolysis method in compost extracts with different dilutions. Non-
diluted - compost extract with no compost inoculum but with ‗fish and kelp hydrolysate‘ and
molasses, 1:10 - compost extract at the ratio of 1 part compost to 10 parts water (w/v), 1:100 -
compost extract at the ratio of 1 part compost to 100 parts water (w/v), 1:1000 - compost extract
at the ratio of 1 part compost to 1000 parts water (w/v). Within a time, treatment means with the
same letter do not differ significantly and values are mean ± SE (n = 3).

There was no difference in the microbial activity of ‗non-diluted‘ (i.e. no

compost) and 1:1000 dilution ratios (Fig. 2.4). In the ‗non-diluted‘ treatment,

oxygen levels were maintained above 80% saturation until 15 h of incubation

period (Fig. 2.3D), after which the oxygen level also declined dramatically. As

the cultures were open, and inputs non sterile, the likely sources of this

contamination are either atmospheric, water or the food source. This ready

contamination of the cultures is indicative of a need to maintain a high compost

to water ratio, to maximise inoculum load.

71
Chapter 2: Compost based extract - characterisation

Compost extracts based on high dilution ratio, e.g. 1:1000, are being

practiced by many farmers in southern Queensland and northern NSW for broad-

acre production (S. Stevens, 2008 pers. comm.). Based on the results of this

study, this practice is not recommended, given the potential for contamination.

Rather, a compost:water ratio of as high as 1:10 is recommended.

3.4. Experiment 3: Comparison of aerated and non-aerated compost

extracts with and without bags

During incubation, significantly higher values of pH and EC were

recorded in aerated compost extracts compared to the non-aerated compost

extracts, irrespective of the use of bags (Fig. 2.5). Although Pant et al. (2009) did

not find any difference in the pH and EC values between aerated and non-aerated

extracts, Scheuerell and Mahaffee (2004) found significantly higher pH and EC

of aerated compost extracts over non-aerated extracts independent of the use of

bacterial or fungal additives.

72
Chapter 2: Compost based extract - characterisation

A 8
7
6
5

pH
4
3
ACE
2 NCE
ACEb
1 NCEb

0
0 3 6 9 12 15 18 21 24 48

B 5.0
4.5
4.0
3.5
EC (dS m-1)

3.0
2.5
2.0
ACE
1.5 NCE
1.0 ACEb
NCEb
0.5
0.0
0 3 6 9 12 15 18 21 24 48

C 32
31
30
29
Temperature (°C)

28
27
26
ACE
25 NCE
24 ACEb
NCEb
23
22
0 3 6 9 12 15 18 21 24 48

D 120
100

80
Dissolved oxygen (%)

60

ACE
40
NCE
ACEb
20 NCEb

0
0 3 6 9 12 15 18 21 24 48
Time (h)

Figure 2. 5. Comparison of pH, electrical conductivity (dS m-1), fluctuations in temperatures (˚C)
and dissolved oxygen (% of saturation) over time (readings taken at every 3 h intervals until 24 h
and final readings taken at 48 h) in rumen compost extracts in presence and absence of aeration
with and without the use of bags. ACE - aerated compost extract with compost kept loose in a bag
while incubating, NCE - non-aerated compost extract with compost kept loose while fermenting,
ACEb - aerated compost extract with compost kept in a bag while incubating, NCEb - non-aerated
compost extract with compost kept in a bag while fermenting. Values are mean ± SE (n = 3).

73
Chapter 2: Compost based extract - characterisation

The level of oxygen in solution was maintained at > 80% saturation until

9-12 h of incubation under an aerated condition, whereas it was at approximately

60% saturation until 3-6 h of incubation in non-aerated treatments where it

declined to close to zero by 12-15 h (Fig. 2.5D). The dissolved oxygen levels of

the non-aerated extracts, lacking extra supply of oxygen, declined 6-9 h earlier

than the aerated extracts. To improve the aeration rate further, an air delivery

system is required that decreases the size of air bubbles, thereby increasing

bubble surface area to volume ratio, and increases the residency time of the air

bubble in the solution (Garcia-Ochoa et al., 2010).

The non aerated cultures were significantly lower in cultivable

microorganisms and their activities when compared to aerated cultures (Table 2.3

and Fig. 2.6), in agreement with the observations of Scheuerell (2004). This

author postulated that anaerobic conditions could result in the production of

reduced organic compounds that are inhibitory to the microbes. Of course,

population growth of microorganisms is greatly influenced by the concentration

of available dissolved oxygen (Metcalf & Eddy, 2003). Tajuddin et al. (2004)

showed an inverse relationship between dissolved oxygen deficit and bacterial

counts when aerating an effluent of palm oil mill for thirteen days. The authors

found that a continuous supply of oxygen in the effluent enhanced the bacterial

population. Kelley (2004) also noted that microbial growth becomes strictly

limited if supplementary oxygen is not provided to the compost extract

incubation system.

74
Chapter 2: Compost based extract - characterisation

Table 2. 3. Comparison of physico-chemical and microbial characteristics of compost extracts


incubated for 24 h in presence and absence of aeration with and without the use of bags.

Parameter ACE NCE ACEb NCEb


+ -1
NH4 -N (µg g d wt) 0.72 ± 0.03b 0.65 ± 0.03b 0.72 ± 0.03b 0.47 ± 0.00a
- -1
PO43 -P (µg g d wt) 34.6 ± 1.00b 36.5 ± 0.38b 26.5 ± 0.58a 25.0 ± 0.22a
+ -1
K -K (µg g d wt) 0.83 ± 0.10a 0.66 ± 0.04a 0.73 ± 0.05a 0.75 ± 0.02a
-1
Humic acid (µg g d wt) 0.59 ± 0.02b 0.59 ± 0.07b 0.61 ± 0.07b 0.23 ± 0.02a
-1
Fulvic acid (µg g d wt) 1.26 ± 0.07a 1.12 ± 0.17a 1.11 ± 0.24a 0.74 ± 0.04a
Bacterial population (cfu Log10) @ 24h 8.62 ± 0.03d 7.62 ± 0.03b 8.28 ± 0.03c 7.40 ± 0.02a
Fungal population (cfu Log10) @ 24h 4.83 ± 0.07b 4.10 ± 0.10a 4.59 ± 0.06b 4.00 ± 0.00a
d wt Dry weight. ACE - aerated compost extract with compost kept loose in a bag while incubating,
NCE - non-aerated compost extract with compost kept loose while fermenting, ACEb - aerated compost
extract with compost kept in a bag, NCEb - non-aerated compost extract with compost kept in a bag. Within
rows, means (n = 3) with the same letter are not significantly different according to Tukey's test (p< 0.05).

Figure 2. 6. Comparison of total microbial activity over time (readings taken at 12, 24 and 48 h)
following the FDA hydrolysis method in compost extracts in presence and absence of aeration
with and without the use of bags. ACE - aerated compost extract with compost kept loose in a bag
while incubating, NCE - non-aerated compost extract with compost kept loose while fermenting,
ACEb - aerated compost extract with compost kept in a bag while incubating, NCEb - non-aerated
compost extract with compost kept in a bag while fermenting. Within a time, treatment means
with the same letter do not differ significantly and values are mean ± SE (n = 3).

Aeration also increased mixing, and thus leaching of solutes and

microbiota from the compost (Table 2.3). In accordance with this, Pant et al.

(2009) found higher macronutrient concentrations of total nitrogen, nitrate, nitrite

and potassium with comparable concentrations of secondary nutrients (Ca, Mg,

75
Chapter 2: Compost based extract - characterisation

Na) and micronutrients (Fe, Mn, Zn, Cu, B) in aerated compost extracts

compared to the non-aerated compost extracts.

The speed at which dissolved oxygen was consumed (Fig. 2.5D) in the

extraction process was greater for the loose compared to the bagged compost

extract. Limited oxygen at the centre of the compost contained in a bag may have

reduced the microbial activity significantly. Although by 12 h there was a

difference in microbial activity between aerated and non-aerated treatments (Fig.

2.6), no difference between loose and bagged compost extract was noted in the

total microbial activity of aerated compost extracts until after 24 h of incubation,

and then there was only an increase in the bagged aerated treatment. In

accordance with the higher oxygen demand (Fig. 2.5), there was a higher

bacterial count (Table 2.3) in non-aerated compost extract without compared to

with the use of bags. These observations reinforces the conclusions of Garcia-

Ochoa et al. (2010) that oxygen or agitation are needed for the survival and

proliferation of microbes in a given liquid culture.

4. Conclusion

Composted cattle paunch material produced extracts dominated by

bacteria, with approximately 2:1 ratio of bacterial to fungal cfu. To achieve

maximum nutrient extraction and microbial population increase, with bias to

fungal growth over bacterial growth, the following conditions are recommended:

(i) addition of both kelp hydrolysate and molasses; (ii) a compost:water ratio of

1:10, to produce microbially enhanced compost extracts rich in soluble

metabolites; (iii) use of bags to avoid clogging of spray nozzles; and (iv) aeration

76
Chapter 2: Compost based extract - characterisation

to maintain bulk solution oxygen levels above 6 ppm. These conditions are used

to produce compost extracts for agronomic trails in companion studies.

77
Chapter 3: Vermiculture based extract - characterisation

Chapter 3

Changes in microbial and nutrient composition

associated with rumen content compost incubation

A version of this chapter has been published in Bioresource Technology by Karuna Shrestha,
Pramod Shrestha, Eric M. Adetutu, Kerry B. Walsh, Keith M. Harrower, Andrew S. Ball and
David J. Midmore

Abstract

Physico-chemical and microbiological investigations were carried out on

rumen content material composted for nine months, fresh vermicasts (obtained

after passing the same compost through the guts of a mixture of three species of

earthworms: Eisenia fetida, Lumbricus rubellus and Perionyx excavates) and

microbially enhanced extracts derived from rumen compost, vermicast and

vermicast leachate incubated for up to 48 h. Compared to composted rumen

contents, vermicast was only improved in terms of microbial biomass C, while

vermicast leached extract was significantly higher in NH4+-N, PO4--P, humic

acid, bacterial counts and total microbial activity compared to rumen compost

extract. Although no difference between treatments was observed in genetic

diversity as indicated by DGGE analysis, community level functional diversity of

vermicast leached extract (BiologTM) was higher than that of composted rumen

78
Chapter 3: Vermiculture based extract - characterisation

contents, vermicast and rumen compost extract indicating an enhancement of

microbial activity rather than diversity due to liquid incubation.

1. Introduction

Composting is a process of aerobic decomposition of organic matter under

controlled conditions by microbiota, producing a stable organic end product, with

carbon-dioxide, water and heat produced as by-products (Rynk et al., 1992). For

large scale composting activity, point sources of organic material of uniform

quality is required. One such material is the partially digested rumen contents of

cattle, viewed as a waste in abattoir operations. For example, the two abattoirs in

Rockhampton slaughter approximately 1700 head a day, producing 140 cubic

meters of rumen content per day (R. Lang, pers. comm.).

Vermicomposting is a process of biotransforming and stabilising organic

materials (often waste) into humus (Neuhauser et al., 1988) by the combined

activity of earthworms and microorganisms (Aira & Dominguez, 2008).

Earthworms excrete partially digested materials (Parmelee et al., 1990), known as

vermicasts or castings, which are more homogeneous in composition than the

source material (Albanell et al., 1988), have reduced levels of contamination

(Ndegwa & Thompson, 2000), and contain elevated levels of plant growth

regulators and/or symbiotic microbes (Kale et al., 1992) and organic acids such as

humic and fulvic acids (Edwards et al., 2006).

There is an emerging commercial trend of aerobically incubating an

extract of compost with a carbohydrate and a protein source, producing a

microbially enhanced liquid (Ingham, 1999a; Pant et al., 2009). Known by the

agricultural sector as ‗compost teas‘ in the current study this microbially

79
Chapter 3: Vermiculture based extract - characterisation

enhanced product is termed ―Compost Extract‖ or ―CE‖ in short. Compost extract

contains nutrients extracted from compost and thus contributes directly to plant

nutrition, and also contains organic matter, improving soil structure and water

holding capacity by building soil aggregates. Compost extracts derived from

vermicomposts have proven benefits in terms of growth and yield promotion of

various agricultural crops (Pant et al., 2009; Tejada et al., 2008). Moreover,

compost extract adds microorganisms to the soil and ‗brings the soil back to life‘

(Martens, 2001). However, at the typically employed rates of application of 150-

200 L/ha, the fundamental aim of compost extract application is to alter soil

microbial populations towards a more plant beneficial community (Hoitink &

Boehm, 1999b).

Many researchers have characterised the composting process (Singh &

Sharma, 2002; Vivas et al., 2009), however, there is little documentation on the

shift in microbial diversity during the incubation and extraction process.

Therefore, a comparative study was undertaken on the effect of vermicomposting

of rumen content compost and of an extract enhancement process. Physico-

chemical, biochemical and microbial characteristics were monitored on the

following: rumen compost extract (RCE), vermicast extract (VCE) and vermicast

leached extract (VLE), i.e. microbially enhanced vermicast leachate collected in

tanks from vermibeds by continuous recirculation. Microbial enhancement was

effected by amendment of a compost, vermicasts or vermicast leachate with

microbial food sources (fish & kelp hydrolysate and molasses) under aerobic

conditions. The null hypothesis in this experiment is that the microbial population

present in the compost is not altered by earthworm feeding, or by the liquid

extract enhancement process.

80
Chapter 3: Vermiculture based extract - characterisation

2. Materials and methods

2.1. Rumen compost

Rumen compost was obtained from a commercial composting operation

(Broadmeadows) which utilises the rumen contents from the Australia Meat

Holding Pty Ltd., Abattoir (maximum daily kill of approximately 1700 cattle) in

Rockhampton, QLD Australia. A windrow composting process was undertaken

over approximately nine months, with windrows turned mechanically at

approximately monthly intervals in order to provide aeration and improve

homogeneity.

Nine month old bulk rumen compost was freshly collected from the

commercial operation to feed the earthworms and for laboratory analysis, several

core samples of which were composited and representative samples were taken.

These compost samples were used immediately or stored in sealed plastic bags at

4 ˚C until use.

2.2. Vermicomposting process, vermicasts and vermicast leachate

A mixture of earthworms (three species; namely, Eisenia fetida,

Lumbricus rubellus and Perionyx excavates) was used in three covered culture

chambers that were supplied from an Australian commercial supplier

(Vermicrobe International Pty LtdTM, Australia, http://www.vermicrobe.com).

These species vary in their adaptation to different climatic conditions, e.g.

Eisenia fetida are hardy and can survive in extreme temperature conditions,

Lumbricus rubellus are well adapted to low temperature (as low as 3-4˚C)

whereas Perionyx excavates are adapted to moderate (e.g. 25˚C) temperature,

which was the reason for using a mixture of three species of earthworms. Mature

81
Chapter 3: Vermiculture based extract - characterisation

earthworms were used to seed each chamber at the rate of 6 kg f wt per chamber

containing approximately 60 kg f wt (0.3 m3) of nine month old aged rumen

compost. Feeding occurred every 7-10 days at approximately 20 kg of composted

rumen paunch per chamber. All of the chambers used in this process were

maintained at ambient temperatures during the whole experimental period

(summer of 2009) in Rockhampton (latitude: 23 22‘ 0.345‖ S and longitude: 150 31‘

0.53‖ E). The vermicompost chambers were irrigated twice each day for 15 min at

10:00 a.m. and 2:00 p.m. with vermicast leachate at 120 L for 15 min to maintain

substrate moisture at around 45-70%. Fresh vermicast was removed after twelve

weeks and representative samples (n = 3) were taken for laboratory analysis.

2.3. Physico-chemical analyses

The pH and electrical conductivity (EC) of the rumen compost and

vermicast samples were measured on 1: 20 (w/v) diluted samples. This dilution

ratio was taken considering the high water absorbance of the tested samples.

Moisture content was calculated after drying at 105˚C for 24 h. Inorganic soluble

NPK (NH4+-N, NO3−-N, PO4−-P and K+-K) were determined on triplicate samples

(d wt) using colorimetric test strips (‗RQflex plus‘, Merck, Germany).

2.4. Preparation of compost extracts and sampling regimes

Compost extracts were prepared from rumen compost (RC) and

vermicasts (VC), with 3 kg f wt of material sealed into a cotton (28 g m-2) bag

and submerged into 30 L tap water in a 60 L plastic bucket, and amended with

1% (v/v) ‗fish & kelp hydrolysate‘ and 0.5% (v/v) molasses at the ratio of 1:10

(w/v). Fish & kelp hydrolysate is an organic product of Searles® which consists

82
Chapter 3: Vermiculture based extract - characterisation

of 78% fish hydrolysate, 20% kelp, 2% stabiliser and 4% fish oil on w/w basis. It

contains 10.3% N, 2.5% P, 2.5% K, 7% Ca, 0.2% Mg, 0.5% S, 690 mg/kg Fe, 5

mg/kg Cu, 14 mg/kg Mn, 32 mg/kg Zn, 14 mg/kg B, and 1 mg/kg Mo in organic

form. Molasses is a treacle-like by-product of cane processing. The water used

was aerated for 30 minutes to remove chlorine before addition of compost.

Compost extracts were incubated indoors at between 22˚C and 30˚C (diurnal

range). The suspensions were continuously aerated (36 L/ min air delivery per

bucket through air stones) for 48 h to produce rumen compost extract and

vermicast extract. Concurrently, same volumes of vermicast leachate (30 L) with

‗fish and kelp hydrolysate‘ and molasses were also incubated in the same manner

producing vermicast leached extract.

The parameters of pH, EC, temperature and dissolved oxygen were

monitored at 3 h intervals for 24 h and finally at 48 h during preparation of the

extracts. Inorganic nutrient analyses of triplicate samples of the compost extracts

was performed using ‗RQflex plus‘ colorimetric test strips. Humic and fulvic

acids were also determined from samples collected at 24 h using a UV-Vis

spectrophotometer following a slightly modified method by Yamada et al. (1998).

FDA hydrolysis was determined using 1 mL samples collected after 12, 24 and

48 h incubation. Microbial populations were determined from 24 h incubated

samples, using the methods discussed below.

2.5. Microbial determinations

Bacterial and fungal populations of serially diluted samples were

determined by the pour plate method (Harrower, 2006). Bacterial plates were

incubated at 37˚C for 1-2 days, while fungal plates were incubated at 25˚C for 3-

83
Chapter 3: Vermiculture based extract - characterisation

5 days, both with three replications. The evolved CO2 of unaltered samples (basal

respiration, BR) and of substrate induced respiration (SIR) was assessed of

triplicate samples as an index of microbial activities of rumen compost and

vermicast samples following the protocol of Anderson (1982). As an alternative

measure of microbial activity, fluorescein diacetate (FDA) hydrolysis was

determined following the method given by Adam and Duncan (2001).

2.6. Microbial biomass carbon analysis

Microbial biomass carbon was determined by the chloroform fumigation

extraction method (Vance et al., 1987). Total and dissolved organic C (TOC and

DOC, respectively) were measured spectrophotometrically following the

dichromate digestion method (Walkley & Black, 1934). The microbial biomass C

was obtained by subtracting DOC from TOC and dividing by a factor used in

converting extracted organic carbon to microbial biomass C, kec, where kec = 0.33

(Sparling & West, 1988).

2.7. Community level physiological profiles of bacteria and fungi

The Ecoplate and FF microplate systems (BiologTM Inc, USA) were used

to identify bacterial and fungal functional diversity, respectively. The rates of

utilisation of 31 different C substrates are assessed in the Ecoplate method, while

the FF microplate involves use of 95 substrates (Insam, 1997). Triplicate samples

of rumen compost and vermicast (1 g f wt) as well as rumen compost extract and

vermicast leached extract samples (1 mL) were serially diluted to 10-3 in 9 mL of

sterile water in triplicate. Solutions were then inoculated to each well of the

Biolog Ecoplates and FF plates, using 100 and 150 μL samples, respectively. The

84
Chapter 3: Vermiculture based extract - characterisation

microplates were monitored at 0, 24, 48, 72, 96, 120, 144 and 168 h by ELISA

Reader at 570 nm for the Ecoplate, and at 490 and 750 nm for the FF microplates.

The average well colour development (AWCD) of all absorbance data was

individually calculated for all different C substrates utilised by the microbiota in

both Ecoplate and FF microplate (Konopka et al., 1998), however, the 72 h

absorbance data were chosen for statistical analysis (Garland, 1996).

2.8. PCR and DGGE

Total DNA was extracted by the bead beating method from three samples

of each preparation type (rumen compost, vermicast, rumen compost extract, and

vermicast leached extract), using a DNA extraction kit (MO BIO Laboratories,

Inc., Carlsbad, CA). Polymerase chain reactions (PCR) were performed using a

thermocycler (Bio-Rad, USA) using 50 µL reaction volumes containing

approximately 5 ng of purified genomic DNA, 0.5 µM of each primer, 2.5 mM

MaCl2, 0.2 mM of dNTP, 10 µL of 5X PCR-buffer and 0.3 U of Taq

Polymerase. The thermocycling conditions for PCR were as described by Girvan

et al. (2003). The PCR products from amplifications of 16S rDNA and ITS

regions were analysed with Universal Mutation Detection System (Bio Rad Inc.,

CA, USA) using 9% polyacrylamide gels (the ratio of acrylamide was 37: 1).

Bacterial and fungal communities were profiled by DGGE analysis of the

appropriate amplicons using a denaturing gradient of 40 to 60% for bacteria and

42 to 52% for fungi, electrophoresed at 60 V for up to 20 h at 60˚C. DGGE gels

were then silver stained (Girvan et al., 2003), scanned and saved as tiff files with

Epson Expression V700 Pro and used for subsequent analysis using Phoretix 1D

advanced analysis package (Phoretix Ltd, UK).

85
Chapter 3: Vermiculture based extract - characterisation

2.9 Statistical analysis

Statistical significance for all replicated samples was determined by

analysis of variance (ANOVA) following Tukey‘s (HSD) pairwise comparison

using the statistical package (XLSTAT, 2010). Differences in means were

compared at p< 0.05. Data on relative band intensities of microbial communities

from the DGGE profile were calculated with Phoretix 1D advanced analysis

package (Phoretix Ltd, UK). The UPGMA dendrogram was generated and the

digitised image was analysed by the same package to express the relatedness of

microbial communities as similarity clusters. The Shannon Weaver diversity

index H‘ (Shannon & Weaver, 1963) and Equitability index J (Begon et al., 1990)

were calculated from the number and intensities of bands present in each lane.

3. Results and discussion

3.1. Physico-chemical analysis

The rumen compost material contained approximately 0.2% w/w N in

soluble form, and 0.05% soluble P (Table 3.1). Vermicast material was lower in

EC, nitrate, phosphate, humic acid and fulvic acid content, compared to the

original compost material. This result is expected, given extraction of nutrients by

the earthworms (which doubled in mass during the vermicomposting period) and

leaching of the vermicast material (1033 L m-3).

86
Chapter 3: Vermiculture based extract - characterisation

Table 3. 1. Initial physico-chemical, microbial and biochemical characteristics of rumen compost


and vermicast (n = 3).

Parameters Rumen compost (RC) Vermicast (VC)


pH (1:20) 6.53 ± 0.02 a 6.62 ± 0.02 a
-1
EC (dS m ) (1:20) 0.35 ± 0.01 b 0.26 ± 0.01 a
Moisture content (%) 58.84 ± 1.51 a 56.56 ± 2.24 a
- -1
NO3 -N (mg kg d wt) 1997.38 ± 29.37 b 1015.25 ± 30.52 a
+ -1
NH4 -N (mg kg d wt) 35.33 ± 2.52 b 21.66 ± 4.17 a
3- -1
PO4 -P (mg kg d wt) 551.46 ± 10.61 b 283.13 ± 10.11 a
+ -1
K -K (mg g d wt) 60.49 ± 11.25 a 50.23 ± 7.23 a
-1
Humic acid (g kg d wt) 45.37 ± 3.14 b 5.67 ± 0.45 a
-1
Fulvic acid (g kg d wt) 21.53 ± 1.56 b 4.54 ± 0.14 a
Bacterial population (cfu Log 10 g-1) 7.88 ± 0.02 b 6.40 ± 0.10 a
Fungal population (cfu Log 10 g-1) 5.19 ± 0.02 b 4.59 ± 0.06 a
-1
Microbial biomass C (mg C g d wt) 3.33 ± 0.45 a 12.56 ± 0.37 b
-1 -1
Basal respiration (µg CO2 g d wt h ) 81.58 ± 3.71 b 38.72 ± 3.52 a
-1 -1
Substrate induced respiration (µg CO2 g d wt h ) 808.39 ± 9.81 b 528.03 ± 3.52 a
-1 -1
Total Microbial activity (µg FDA g d wt h ) 1218.98 ± 176.95 b 611.34 ± 41.30 a
d wt Dry weight
Within rows, means with the same letter are not significantly different according to Paired
T- test (p< 0.05).

3.2. Microbial and biochemical assays

Plate counts of bacteria and fungi, and basal and substrate induced

respiration were significantly higher for the initial rumen compost than for

vermicast, however, microbial biomass was lower (p< 0.0001) and extractable

DNA (data not shown) was not different between the two substrates (Table 3.1).

These results are consistent with a more 'active' microflora in the compost than

the vermicast, the latter in which microbiota were apparently dormant.

3.3. Characterisation of compost extracts

Solid compost materials are bulky to apply in broad scale

horticulture/agriculture, and thus there is a growing interest in liquid extracts or

microbially enhanced liquids based on incubation of source inoculum. In this

study, the initial compost, the vermicast material and the vermicast leachate were

87
Chapter 3: Vermiculture based extract - characterisation

used as source inocula; all of which were amended with the ‗fish and kelp

hydrolysate‘ and molasses, producing rumen compost extract, vermicast extract

and vermicast leached extract, respectively.

During incubation significantly higher pH values and temperature were

recorded in vermicast leached extract, while rumen compost extract had

significantly higher EC compared to other liquid extracts (Fig. 3.1). The oxygen

level in vermicast extract was maintained at approximately 80% saturation until 9

h of incubation period. However, dissolved oxygen decreased rapidly after

incubation for 3-6 h in vermicast leached extract, 6-12 h in rumen compost

extract, and 9-12 h in vermicast extract, indicative of the level of microbial

activity during incubation (Fig. 3.1D). The depletion in oxygen recorded during

incubation indicates that the aeration system was not capable of meeting the

oxygen demand of the microbial culture. This could be addressed by either

improving the aeration system (e.g. small bubbles, increased air flow rate) or by

decreasing microbial demand (e.g. provision of less substrate). Such comparative

studies on microbially enhanced compost extracts produced with different levels

of microbial demand have been studied in another experiment (data not

presented).

88
Chapter 3: Vermiculture based extract - characterisation

Figure 3. 1. Comparison of pH, EC (dS m-1), fluctuations in temperatures (°C) and dissolved
oxygen (%) per unit time (readings taken at every 3 h intervals until 24 h after commencement of
compost extract incubation and final readings taken at 48 h) in rumen compost extract (RCE),
vermicast leached extract (VLE) and vermicast extract (VCE) (n = 3).

89
Chapter 3: Vermiculture based extract - characterisation

The bulk solution oxygen level was < 5% for 6-21 h in vermicast leached

extract, for12-24 h in vermicast extract and for 12-48 h in rumen compost extract,

despite the fact that these solutions were being aerated continuously at the rate of

36 L air/min per 30 L vessel. Given non-ideal mixing and aeration, the culture is

likely to have consisted of a mixture of environments ranging from aerobic to

anaerobic. Such conditions may favour increased microbial diversity.

Vermicast leached extract was high in ammonium, phosphate and humic

acid, relative to the rumen compost extract (Table 3.2), as expected considering

the metabolising activity of earthworms. Potassium and fulvic acid levels were in

the order of rumen compost extract > vermicast leached extract > vermicast

extract, and rumen compost extract = vermicast leached extract > vermicast

extract, respectively. Bacterial and fungal populations as indicated by the plate

counts were in the order vermicast extract > vermicast leached extract > rumen

compost extract and vermicast extract > rumen compost extract > vermicast

leached extract, respectively. In contrary to the pour plate results, highest

microbial activity in vermicast leached extract and least in vermicast extract was

demonstrated by the FDA hydrolysis method (Fig. 3.2). The difference between

the two methods presumably reflects the enumeration of culturable organisms

under the conditions of the pour plate, while the FDA method assesses all active

organisms in the culture at time of assay.

90
Chapter 3: Vermiculture based extract - characterisation

Table 3. 2.Comparison of chemical, biochemical and microbial characteristics of rumen compost


extract (RCE), vermicast leached extract (VLE), vermicast extract (VCE) incubated for 24 hours
with 1% kelp & fish hydrolysate and 0.5% molasses as amendments in all treatments (n = 3).

Parameters RCE VLE VCE


+ -1
NH4 -N (mg L ) 1.06 ± 0.05 b 1.42 ± 0.03 c 0.23 ± 0.04 a
3- -1 21.09 ± 0.87 a 26.31 ± 1.15 b 30.22 ± 1.21 b
PO4 -P (mg L )
+ -1
K -K (g L ) 0.68 ± 0.01 b 0.60 ± 0.02 b 0.46 ± 0.02 a
-1
Humic acid (g L ) 0.70 ± 0.11 b 2.23 ± 0.13 c 0.04 ± 0.01 a
-1
Fulvic acid (g L ) 1.12 ± 0.12 b 0.84 ± 0.03 b 0.33 ± 0.02 a
Bacterial population (cfu Log 10 mL-1) 11.50 ± 0.03 a 11.88 ± 0.01 b 11.99 ± 0.01 c
Fungal population (cfu Log 10 mL-1) 5.16 ± 0.04 b 4.89 ± 0.08 a 5.64 ± 0.03 c
d wt Dry weight
Within rows, means with the same letter are not significantly different according to
Tukey's test (p< 0.01).

Figure 3. 2. Comparison of total microbial activity per unit time (readings taken at 12, 24 and 48
h) following the FDA hydrolysis method in rumen compost extract (RCE), vermicast leached
extract (VLE) and vermicast extract (VCE) (n = 3).

3.4. Community level physiological profiles of bacteria and fungi

The functional diversity of culturable microorganisms of the tested

samples were differentiated on the basis of the bacterial and fungal kinetics and

substrate utilisation patterns (Konopka et al., 1998). Although total microbial

activity (as measured with FDA) in rumen compost was twice that of vermicast

(Table 3.1), no difference was observed between rumen compost and vermicast in

91
Chapter 3: Vermiculture based extract - characterisation

terms of the rate of substrate utilisation by bacteria (BiologTM Ecoplate), substrate

utilisation by fungi (FF microplate) was significantly higher in rumen compost

compared to the vermicast (Fig. 3.3) after 72 h of incubation. However, a sudden

decline in optical density (AWCD 490 nm) of rumen compost was noted in the

FF microplate readings at 120 h compared to that measured at 96 h of incubation.

The reason for this could be due to the antagonistic interactions between the

microbiota or due to the toxicity caused to the microorganisms by the redox dyes

used in the Biolog wells as suggested by Ullrich et al. (1996) who studied the

toxic effects of redox dye on bacterial metabolism. However, the exact reason for

this temporary decline is unknown. Biolog plate substrate utilisation by vermicast

leached extract was significantly higher than that by rumen compost extract (Fig.

3.3), consistent with the results obtained from the FDA hydrolysis (Fig. 3.2).

Thus, the incubation of the compost extracts with additional food sources allowed

an increase in the active bacterial and fungal population compared to that of the

original communities of the composts as suggested by other researchers (Naidu et

al., 2010; Scheuerell, 2004).

92
Chapter 3: Vermiculture based extract - characterisation

Figure 3. 3. Kinetics of average well colour development (AWCD) and optical densities of (A)
bacterial (570 nm) and (B) fungal (490 nm) communities, originating from rumen compost (RC)
and vermicast (VC) as well as rumen compost extract (RCE) and vermicast leached extract (VLE)
(n = 3).

Principal component analysis of the Biolog data indicated that rumen

compost and vermicast were different in bacterial and fungal substrate utilisation

patterns compared to rumen compost extract and vermicast leached extract (with

separation by PC1, Fig. 3.4). The PCA of Ecoplate data explained 91% (PC1) and

5% (PC2) of the total variance whereas the PCA of FF microplate data explained

84% (PC1) and 7% (PC2) of the total variance (Fig. 3.4). In the present study, we

did not include a statistical control, i.e. rumen compost without earthworms,

because the aim of this experiment was to generate vermicast, which is compost

93
Chapter 3: Vermiculture based extract - characterisation

that has passed through the gut of the earthworm and compare that with the initial

compost. It could be that the changes observed in the nutrient and microbial

composition of the vermicast could have been related to the duration of the

incubation and passage through the earthworm. However, substrate utilisation

results cannot confirm the differences in species composition of microbiota as

only the functional physiological profile of culturable microorganisms can be

detected by this method (Widmer et al., 2001).

Figure 3. 4. Principal component analysis of optical densities produced by microbial activities of


rumen compost (RC), vermicast (VC), rumen compost extract (RCE) and vermicast leached
extract (VLE) consuming 32 (Biolog Ecoplate) and 96 (Biolog FF microplate) different food
sources including the control representing (A) bacterial and (B) fungal diversity, respectively (n =
3).

94
Chapter 3: Vermiculture based extract - characterisation

3.5. PCR-DGGE analysis

Since it is based on extractable DNA, the PCR-DGGE method should

represent an assessment of microbial diversity, however, DGGE will only detect

the dominant 1% of species in the bacterial population (Murray et al., 1996).

Derived measures of H‘ (diversity), J (evenness) and S (richness) from DGGE gel

pictures were similar for all tested materials (rumen compost, vermicast, rumen

compost extract and vermicast extract), for both bacteria and fungi (Table 3.3).

The only difference was a low result for bacterial S values in rumen compost.

These results suggest a similar species range of microorganisms in all products.

The results of PCR-DGGE analysis were consistent with that of the Biolog, in

that the bacterial community structure of the rumen compost and vermicast were

more similar to each other than to the liquid cultures (Fig. 3.5A). The UPGMA

dendrogram showed 45% homology between all the bacterial communities, while

55% and 57% homology was seen between the solid products (rumen compost

and vermicast) and the liquid extracts (rumen compost extract and vermicast

extract), respectively (Fig. 3.5A). The difference between replicate batches was

low in some cases (e.g. vermicast leached extract and vermicast), but high in

others (e.g. rumen compost) (Fig. 3.5A). Similar results were recorded for the

fungal community analysis (Fig. 3.5B), although amplifiable DNA was

unfortunately not obtained from vermicast leached extract samples.

95
Chapter 3: Vermiculture based extract - characterisation

Table 3. 3. Shannon weaver Diversity Index (H‘), Equitability Index (J) and number of bands (S)
for bacteria and fungi present in rumen compost (RC), vermicast (VC), rumen compost extract
(RCE) and vermicast leached extract (VLE) (n = 2).

Bacteria Fungi
Treatments H' J S H' J S
RC 2.42 ± 0.09 a 0.81 ± 0.02 a 20 ± 1.00 a 2.90 ± 0.02 a 0.90 ± 0.01 a 25 ± 0.00 a
VC 2.49 ± 0.02 a 0.76 ± 0.01 a 26 ± 0.00 ab 2.98 ± 0.02 a 0.90 ± 0.00 a 27 ± 0.00 a
RCE 2.51 ± 0.22 a 0.78 ± 0.05 a 25 ± 2.00 ab 2.60 ± 0.28 a 0.87 ± 0.05 a 20 ± 3.00 a
VLE 2.86 ± 0.00 a 0.83 ± 0.00 a 31 ± 0.00 b ND ND ND
ND not determined
Within rows, means with the same letter are not significantly different according to Tukey's test (p< 0.05).

Figure 3. 5. UPGMA (Unweighted Pair Group Method with Arithmetic mean algorithm)
dendrogram generated from responses to 16S rDNA and Internal Transcribed Spacer region based
Denaturing Gradient Gel Electrophoresis (DGGE) analyses for (A) bacteria and (B) fungi present
in rumen compost (RC), vermicast (VC), rumen compost extract (RCE) and vermicast leached
extract (VLE) (n = 2).

Higher functional diversity in vermicast was found compared to the

original compost as revealed by the Shannon-Weaver diversity index H‘ (Sen &

Chandra, 2009). Similarly, Vivas et al. (2009) showed that finished vermicast

96
Chapter 3: Vermiculture based extract - characterisation

was richer in bacterial population and functional diversity than the original and

composted olive-mill waste, with increased values of H‘ calculated using DGGE

banding patterns. In contrast, rumen compost extract and vermicast leached

extract were not higher in diversity index (H‘) or equitability index (J) for

bacteria or fungi than the original rumen compost or vermicast although greater

number of bands in the bacterial DGGE profile (band richness, S) was observed

in vermicast leached extract, relative to the original rumen compost (Table 3.3).

This indicates that there was no significant difference in the bacterial or fungal

communities of either solid or liquid phases of rumen compost and vermicast. It

may be that earthworm action results in little change in mirobial species diversity

for well composted material, in comparison to incompletely composted material.

The olive mill waste of Vivas et al. (2009) may have fallen into the latter

category.

4. Conclusions

Coupled with chemical, biochemical and microbial approaches, rumen

compost before and after passing through the guts of earthworms was evaluated

and compared with their subsequent microbially enhanced extracts. While

earthworm activity increased the microbial biomass of the rumen compost,

liquid incubation of composts altered their microbial and nutrient composition.

Although no differences were found in the genetic diversity of microorganisms,

highest functional diversity was observed in vermicast leached extract compared

to rumen composts, vermicast and rumen compost extract. The ability of these

liquid extracts to shift the microbial activity and diversity of the soil is another

fertile area for consideration.

97
Chapter 4: Commercial extract - characterisation

Chapter 4

Comparison of microbially enhanced compost

extracts produced from composted rumen content

material and from commercially available inocula

A version of this chapter has been published in Bioresource Technology by Karuna Shrestha, Eric
M. Adetutu, Pramod Shrestha, Kerry B. Walsh, Keith M. Harrower, Andrew S. Ball and David J.
Midmore

Abstract

A comparative study was performed on compost extracts prepared from

cattle rumen content composted for three and nine months, nine month old

compost inoculated with a Nutri-Life 4/20TM inoculum, and two commercial

preparations (Living SoilTM and Nutri-Life 4/20TM), all incubated for 48 h. Nutri-

Life 4/20TM had the highest concentrations of NO3−-N and K+-K, while rumen

compost extract had higher humic and fulvic acids concentration. The bacterial

and fungal community level functional diversity of three month old compost

extract and of Living SoilTM, assessed with BiologTM, were higher than that of

nine month old rumen compost extract, with or without Nutri-Life 4/20TM

inoculum, or Nutri-Life 4/20TM. No difference in fungal diversity was observed

between treatments, as indicated by Denaturing Gradient Gel Electrophoresis

(DGGE) analysis, however, bacterial diversity was higher in all compost extracts

98
Chapter 4: Commercial extract - characterisation

and LivingSoilTM, compared to the Nutri-Life 4/20TM. Criteria for judging the

quality of a microbially enhanced extract are discussed.

1. Introduction

There is an increasing global interest in the use of liquid organic fertilisers

and inoculums of beneficial microbes in support of ‗biological farming‘ and

sustainable agriculture (Naidu et al., 2010). A range of benefits are claimed for

these preparations, including direct nutrient addition, increased mineralisation

and solubilisation, protection against plant pathogens, and mitigation of soil

aluminium toxicity. As an example of an inoculum, compost extracts can be

incubated, either aerobically or anaerobically, with microbial food sources to

create a rich, if uncontrolled, blend of microorganisms. Quite clearly, the quality

of the compost (in terms of minerals and microbiota) will most likely influence

the quality of such extracts. Alternatively, more defined microbial populations

may be used. A number of commercial suppliers of inocula are attempting to

serve this market, typically on a relatively small and local scale. The commercial

products claim to be more consistent than that based on compost, and capable of

being stored (and thus transported). With the increasing use of such amendments

comes a need to establish their benefits, and to set quality criteria.

LivingSoilTM (abbreviated as ―LS‖) (O‘Grady Rural, Lismore, NSW) and

Nutri-Life 4/20TM (abbreviated as ―NL-4/20‖) (Nutri-Tech Solutions, Yandina,

Qld) are examples of microbial formulations that are produced commercially in

Australia. LivingSoilTM is a dry formulation, claimed to contain a blend of

microbiota, plant hormones and humic acids derived from composts or

vermicomposts, with diatomaceous earth acting as a microbial carrier and

99
Chapter 4: Commercial extract - characterisation

microbial food (baker‘s yeast) supplied separately. It is reported to be based on

pure cultures of a number of microorganisms, along with compost and

vermicompost, with the final product claimed to contain Azotobacter spp.,

Azospirillum brasilense, Bacillus spp., Cellulomonas cellosea, Chaetomium spp.,

Metarhizium anisopliae, Pseudomonas spp., Polyanguim cellulosum,

Streptomyces spp., Trichoderma spp., together with unidentified microbes from

the compost and vermicompost (www.smartbugs.com.au). These microorganisms

are involved in nitrogen fixation (e.g. Azotobacter spp. and Azospirillum spp.,

Khammas & Kaiser, 1992), phosphate solubilisation and nutrient release (e.g.

Azotobacter spp., Garg et al., 2001), pectin decomposition (e.g. Bacillus spp.,

Halsall & Gibson, 1985), disease suppression/protection against plant pathogens

(e.g. Trichoderma spp., Harman, 2006), soil bioremediation (e.g. Trametes spp.

Atagana, 2009), and cellulose degradation (e.g. Cellulomonas spp., An et al.,

2005). The inoculum is intended to be used to seed a non-sterile culture, with two

food sources (molasses and a solid powder based on yeast). Thus, this product

emulates a ‗compost tea‘, with the LivingSoilTM product replacing fresh compost

as the inoculum source.

Nutri-Life 4/20TM, sourced from Nutri-Tech Solutions Pty Ltd (Yandina,

Qld. Australia, www.nutri-tech.com.au), is a blend of microbes, and is claimed to

contain Trichoderma spp., Rhizobium spp., Bacillus spp. and Pseudomonas spp.

Nutri-Life 4/20TM is recommended for direct application to the soil at the rate of

30 L ha-1, for addition to compost, and to recover previously disease-affected

soils. Nutri-Life 4/20TM is produced by incubating 1 L of Liquid Microbe

FoodTM, 1 L of Dominate (fungi)™ and 50 g of Nutri-Life 4/20TM inoculum per

100 L of non-chlorinated water. Dominate (fungi)™ is claimed to create

100
Chapter 4: Commercial extract - characterisation

favourable conditions for a fungal dominated brew. Liquid Microbe FoodTM is

described as a fungal and bacterial food source, and Nutri-Life 4/20TM inoculum

as a freeze-dried formulation of bacterial and fungal source, enriched in

Trichoderma.

Given the claimed benefits of compost extracts and commercial

biofertilisers, the characterisation of such products has become an interesting area

of investigation. Scheuerell (2004) commented on the need to explore the impact

of production process on the abundance, diversity and evenness of

microorganisms in the resultant compost extracts using molecular methods.

Indeed, Soil Foodweb Institute (SFI) Pty Ltd has operated since 2001 in Australia

providing a microscopically based assessment service. Other techniques used in

this study also have the potential to characterise compost extracts. The present

study aimed to compare microbially enhanced compost extracts produced from

cattle rumen content materials (three and nine month old compost, and nine

month compost along with Nutri-Life 4/20TM inoculum) and two commercial

products (LivingSoilTM and Nutri-Life 4/20TM). The chemical, biochemical and

microbial parameters of the given formulations were compared during incubation

using the same techniques as described in Shrestha et al. (2011a).

2. Materials and methods

2.1. Source compost and commercial preparations

Cattle rumen content material, obtained from Broadmeadows Pty Ltd,

Emu Park Road, Australia, was composted following a windrow composting

process over nine months. Those windrows were turned mechanically at monthly

intervals to provide aeration and to improve homogeneity. Rumen compost aged

101
Chapter 4: Commercial extract - characterisation

three and nine months, achieved from the same source material, was chosen for

this study and characterised (data not shown here). LivingSoilTM was obtained

from O‘Grady Rural (courtesy of Ray O‘Grady, Lismore, NSW) whereas Nutri-

Life 4/20TM was supplied by Nutri-Tech Solutions Pty Ltd, Yandina, Qld

(courtesy of Graeme Sait, NTS).

2.2. Preparation of compost extracts and commercial extracts

Three and nine month old rumen compost extracts were prepared by

completely submerging the respective composts in tied cotton bags, in the ratio of

1:10 (w/v) in tap water amended with 1% (v/v) ‗fish and kelp hydrolysate‘ and

0.5% (v/v) molasses as described in Shrestha et al. (2011a). LivingSoilTM, a

microbially enhanced solution without the use of compost, was produced by

incubating 450 g of LivingSoil inoculant, 150 g MicroBooster CN30 (microbial

food, revealed to contain yeast cells by microscopic examination) (as supplied by

O‘Grady Rural) and 300 mL of molasses, with an addition of 3 mL of a surfactant

(Oresome Antifoam), per 30 L of water, following the manufacturer‘s directions.

Similarly, Nutri-Life 4/20TM was also used to produce a microbially enhanced

solution, following the recommended incubation instructions (300 mL Liquid

Microbe FoodTM) with 300 mL DominateTM and 15 g Nutri-Life 4/20TM inoculum

per 30 L of culture.

A nine month old compost extract was also amended at the beginning of

the incubation with Nutri-Life 4/20TM inoculum (abbreviated as ‗9MTCE‘) at the

rate of 15 g per 30 L of extract. Nutri-Life 4/20TM inoculum is not marketed as a

compost extract amendment. However, we postulated that it could serve as an

extract inoculum, introducing groups such as Trichoderma.

102
Chapter 4: Commercial extract - characterisation

In all the preparations, water was aerated for 30 min for dechlorination

prior to use, and all compost extracts were aerated continuously throughout

incubation with two air stones suspended at the bottom of each bucket, supplying

36 L min-1 of air per 30 L of liquid culture as described in Shrestha et al. (2011a).

All of the preparations were incubated in triplicate for the same duration, up to a

maximum of 48 h, for comparison purpose.

2.3. Physico-chemical analyses

The pH, electrical conductivity (EC), dissolved oxygen, and temperature

of the liquid cultures were monitored at every 3 h interval from 0 to 24 h of

incubation, with an additional measure taken at 48 h. Samples of each culture

were collected in triplicates at 24 h of incubation and concentrations of NH4+-N,

PO4−-P were measured using colorimetric test strips (RQflex plus). For humic

and fulvic acids, samples were centrifuged and filtered through a 0.45 µm

membrane filter and were measured following a slightly modified method by

Yamada et al. (1998).

2.4. Biochemical and microbial analyses

Biological activity in compost extract samples collected at 0, 12, 24, and

48 h was estimated following the FDA hydrolysis method, while microbial

populations were enumerated from samples collected at 24 and 48 h of incubation

using the pour plate count method (Harrower, 2006). For comparative molecular

analyses between treatments, the DNA of the 24 h samples was extracted as

described in Shrestha et al. (2011a) and stored at -20˚C for further use.

103
Chapter 4: Commercial extract - characterisation

2.5. Community level physiological profiles

The Ecoplate and FF microplate systems (Biolog Inc, USA) were used to

identify bacterial and fungal functional diversity of the triplicate samples from all

treatments, respectively. The average well colour development (AWCD) was

calculated at 24 h intervals from 0 to 168 h for all different C substrates in both

Ecoplate and FF microplate (Konopka et al., 1998) before subjecting the 72 h

data to statistical analysis (Garland, 1996).

2.6. Microscopic analyses

Samples of LivingSoilTM and nine month old rumen compost extract were

shipped in cooled liquid form (i.e. with ice blocks) to the Soil Foodweb Institute

(Lismore, NSW) for microbial analysis, where samples were analysed using a

microscopic method. Active and total counts of bacteria and fungi, their

respective ratios, total nematodes, protozoa, hyphal diameter of fungi, and plant

available N supply were the major attributes taken into consideration.

2.7. Genetic diversity of microbial communities of compost extracts

For DNA fingerprinting of rumen and commercial cultures, the 16S

rDNA and Internal Transcribed Spacer region (ITS) based DGGE profiles were

analysed. In this process, DNA of respective samples was extracted in duplicate

and PCR-amplified using universal primers for bacteria and fungi separately. The

amplicons (20 µL of PCR products) for both bacterial and fungal PCR were

subjected to polyacrylamide gel electrophoresis using urea formamide (40 to 60%

for bacteria and 42 to 53% for fungi) as the lower and higher denaturant,

respectively (Girvan et al., 2003), and run for 20 h at 60˚C and 60 V for bacteria

104
Chapter 4: Commercial extract - characterisation

and for 15 h at 58˚C and 78 V in case of fungi using the Bio-Rad DCode system

(Bio Rad Inc., CA, USA). The UPGMA dendrograms were generated and the

Shannon Weaver diversity index (H‘) (Shannon & Weaver, 1963), equitability

index (J) (Begon et al., 1990) and species richness (S) (Krebs, 1999) were then

calculated.

2.8. Statistical analysis

Statistical analyses for all data including values for microbial diversity,

equitability and richness were performed by one or two way ANOVA using a

statistical package (XLSTAT, 2010). Differences between means were assessed

by Tukey‘s (HSD) pairwise comparison at p< 0.05 (Shrestha et al., 2011a).

3. Results and discussion

3.1. Culture characterisation

The Nutri-Life 4/20TM based culture had two orders of magnitude higher

concentrations of NO3−-N, suggesting the presence of a nitrate salt in this product,

while the compost based products had approximately twice the level of PO4−-P,

humic and fulvic acids than the commercial preparations (Table 4.1). The latter

result is expected as rumen based compost used in this study was rich in

phosphate and organic acids, as presented in Shrestha et al. (2011a). There was

little difference in ammonium levels between treatments.

Cultures based on Living soilTM possessed a high pH, around 8, while

those based on three month old rumen compost extract began at pH 6 and

increased to 8 within 48 h (Fig. 4.1). These observations are consistent with the

presence of alkaline material in the Living soilTM formulation, while the increase

105
Chapter 4: Commercial extract - characterisation

in pH with incubation presumably reflects uptake of more anions than cations by

the culture. The pH of nine month old rumen compost extract plus Nutri-Life

4/20TM inoculum decreased from 5.5 to 5 at 48 h (Fig. 4.1), the reason for which

is unclear. It is, however, reported that neutral pH is optimum for the survival of

most microbes isolated from compost (Adegunloye et al., 2007). Future work

should consider pH adjustment of cultures, maintaining a pH of around 6.5.

Table 4.1. Comparison of physico-chemical and microbial characteristics of different liquid


cultures sampled at 24 h and at 48 h of incubation. 3MCE - three month old rumen compost
extract, 9MCE - nine month old rumen compost extract, 9MTCE - nine month old rumen compost
extract plus Nutri-Life 4/20TM inoculum, LS - Living soilTM, NL-4/20 - Nutri-Life 4/20TM. Within
rows, means (n = 3) with the same letter are not significantly different according to Tukey‘s test
(p< 0.05).
Parameter 3MCE 9MCE 9MTCE LS NL-4/20
- -1
NO3 -N (mg L ) 1.20 ± 0.88a 3.61 ± 0.60a 2.56 ± 0.75a 1.13 ± 0.00a 223 ± 14.17b
+ -1
NH4 -N (mg L ) 0.83 ± 0.03b 0.36 ± 0.03a 0.41 ± 0.05a 0.54 ± 0.04ab 0.54 ± 0.13ab
- -1
PO43 -P (mg L ) 25.0 ± 2.69c 14.8 ± 1.25b 25.0 ± 0.78c 6.74 ± 1.39a 5.33 ± 0.58a
+ -1
K -K (g L ) 0.74 ± 0.03a 0.79 ± 0.03a 0.80 ± 0.03a 0.84 ± 0.07a 1.09 ± 0.06b
-1
Humic acid (g kg d wt) 1.31 ± 0.01c 0.61 ± 0.08b 0.49 ± 0.13b 0.06 ± 0.01a 0.03 ± 0.01a
-1
Fulvic acid (g kg d wt) 1.54 ± 0.01c 1.43 ± 0.02c 1.46 ± 0.05c 1.03 ± 0.10b 0.13 ± 0.03a
Bacterial population (cfu Log 10) @ 24h 7.90 ± 0.03c 6.00 ± 0.00a 6.00 ± 0.00a 7.21 ± 0.05b 6.10 ± 0.10a
Bacterial population (cfu Log 10) @ 48h 8.98 ± 0.00c 8.41 ± 0.04ab 8.53 ± 0.04b 8.50 ± 0.06b 8.22 ± 0.05a
Fungal population (cfu Log 10) @ 24h 5.18 ± 0.05a 5.95 ± 0.03bc 5.93 ± 0.04b 6.11 ± 0.04c 6.40 ± 0.01d
Fungal population (cfu Log 10) @ 48h 6.81 ± 0.03a 7.14 ± 0.02b 6.81 ± 0.01a 6.74 ± 0.03a 7.27 ± 0.02c
d wt Dry weight

106
Chapter 4: Commercial extract - characterisation

Figure 4.1. Comparison of (A) pH, (B) electrical conductivity (dS m-1), (C) temperatures (˚C) and
(D) dissolved oxygen (ppm) over time (readings taken at every 3 h intervals until 24 h and final
readings taken at 48 h) in different cultures. 3MCE - three month old rumen compost extract,
9MCE - nine month old rumen compost extract, 9MTCE - nine month old rumen compost extract
plus Nutri-Life 4/20TM inoculum, LS - Living soilTM, NL-4/20 - Nutri-Life 4/20TM (n = 3).

Garcia et al. (1991) reported an increase in EC during composting,

presumably representing a release of mineral ions as organic matter decays.

Although Living soilTM started with the lowest EC, it increased with time in the

Living soilTM incubation (Fig. 4.1), which most likely represents solubilisation or

mineralisation of substrate. The EC of other extracts were relatively stable, after

107
Chapter 4: Commercial extract - characterisation

an increase during the first 3 h of incubation. This observation is consistent with

an extraction of soluble salts from the compost within 3 h, despite the

containment of the compost in a mesh bag. The final extract EC values of around

4 dS m-1 are well within the range tolerated by most plants, indicating that plants

could be grown on neat extracts (‗organic hydroponics‘).

Solution temperature was not significantly different between treatments

(data not shown), except in the last measurement period, at which point three

month old rumen compost extract and Living soilTM treatments were of higher

temperature, suggesting higher microbial growth and activity than present in

Nutri-Life 4/20TM, nine month old rumen compost extract, and nine month old

rumen compost extract plus Nutri-Life 4/20TM inoculum.

A sharp decline in solution dissolved oxygen, reflecting a microbial

oxygen demand greater than the oxygen transfer rate into the solution, occurred

in the three month old rumen compost extract treatment after 6 h of incubation,

after 9 h in the Living soilTM treatment, and later in other treatments (Fig. 4.1D).

This oxygen demand will parallel microbial growth as suggested by Shrestha et

al. (2011a). Here, although equal amount of oxygen was supplied in each

treatment, the level of dissolved oxygen varied among them indicating variation

in microbial population and their activities. The delayed start of microbial activity

in nine month old rumen compost extract, nine month old rumen compost extract

plus Nutri-Life 4/20TM inoculum and Nutri-Life 4/20TM could be due to lower

density of microorganisms in those treatments. The food supply must have

sustained their growth and activity until 24 h of incubation after which the

dissolved oxygen dropped significantly.

108
Chapter 4: Commercial extract - characterisation

3.2. Microbial and biochemical analyses

Plate counts of bacteria were significantly higher in three month old

rumen compost extract, whereas fungal counts were significantly higher in Nutri-

Life 4/20TM samples, whether collected at 24 and 48 h of incubation (Table 4.1).

Higher bacterial counts in three month old rumen compost extract compared to

nine month old rumen compost extract could be due to the aging of source

compost during which the readily available fractions of carbon gets depleted

(Garcia-Gomez et al., 2003; Wu and Ma, 2002). The addition of beneficial fungi

and microbial foods favouring fungal growth, such as Liquid Microbe FoodTM,

Dominate (fungi)TM, and Nutri-Life 4/20TM inoculum, must be the reason for

higher fungal counts in Nutri-Life 4/20TM treatment. Adam and Duncan (2001)

reported a potential bias of the FDA hydrolysis method towards enzymatic

activities of bacteria rather than that of fungi, and total microbial activity as

measured by the FDA hydrolysis was highest by far in three month old rumen

compost extract, followed by that of Living soilTM at 12, 24 and 48 h of

incubation, compared to other treatments (Fig. 4.2). However, the oxygen

demand of the three month old rumen compost extract culture was also higher

than that of any other treatment (Fig. 4.1). Higher microbial activity in ‗young‘

(three month) compost (i.e. higher plate counts and FDA hydrolysis) is expected,

with a decrease in extractable organic carbon during the latter curing phase of

composting, as demonstrated by Wu and Ma (2002). The reason behind low

values in nine month old rumen compost extract and nine month old rumen

compost extract plus Nutri-Life 4/20TM inoculum compared to three month old

rumen compost extract could be due to aging of the compost. The aged compost

must have low organic component compared to the young compost. However,

109
Chapter 4: Commercial extract - characterisation

after 48 h of incubation, the density of microorganisms (Table 4.1) as well as

microbial activity (Fig. 4.2) increased dramatically. This must have resulted in

the large oxygen consumption at 48 h of incubation in the respective cultures as

shown in Fig. 4.1.

Figure 4.2. Comparison of total microbial activity at different times (readings taken at 12, 24 and
48 h) following the FDA hydrolysis method in different cultures. 3MCE - three month old rumen
compost extract, 9MCE - nine month old rumen compost extract, 9MTCE - nine month old rumen
compost extract plus Nutri-Life 4/20TM inoculum, LS - Living soilTM, NL-4/20 - Nutri-Life 4/20TM
(n = 3). Within a time, treatment means with the same letter do not differ significantly and values
are mean ± SE (n = 3).

The addition of Nutri-Life 4/20TM inoculum to the nine month compost

did not result in higher bacterial or fungal plate counts nor enzymatic activities of

microbiota compared to the extract based on nine month compost alone (Fig.

4.2). The plateau population can be expected to be limited by food supply, not

inoculums. Certainly the food sources provided to the culture will influence

microbial species composition. For example, Naidu et al. (2010) reported that

humic acid and yeast extract in the ratio of 4:7 w/w, added as 1 g per 100 g of

compost, enhanced microbial growth (total bacteria, fungi and actinomycetes

measured as cfu mL-1 using spread plate method) during incubation of compost

110
Chapter 4: Commercial extract - characterisation

extracts (prepared at the dilution of 1:5 w/v compost to water). The authors found

that humic acid favoured fungal (e.g. Trichoderma spp., yeast, actinomycetes and

other filamentous fungi) growth up to 61%, while the yeast extract contributed to

bacterial proliferation (e.g. Pseudomonas spp., lactic acid bacteria). There was no

evidence of a quicker population growth in the Nutri-Life 4/20TM inoculum

amended culture. Presumably microbes present in the additive were outcompeted

by those in the compost.

3.3. Community level physiological profiles of bacteria and fungi

Although total microbial activity of three month old rumen compost

extract was the highest amongst all extracts from 12 h onwards (Fig. 4.2),

bacterial communities of Living soilTM showed equal functional diversity to that

of three month old rumen compost extract (Fig. 4.3) as indicated by their similar

patterns and non-significant differences in AWCD values at 72 h of incubation in

wells (Fig. 4.3). The pattern of microbial activity represented by the FDA was

consistent with the pattern of activity indicated by the FF microplate analysis.

The Nutri-Life 4/20TM treatment showed the lowest metabolic rates of substrate

utilisation of both bacteria and fungi (Fig. 4.3). This is consistent with the

bacterial plate counts but inconsistent with the results of fungal plate counts with

high fungal cfu recorded (Table 4.1). The single medium used in the Petri plates

may have favoured a few types of fast growing fungi, compared to the FF

microplate (which contains 95 different carbon substrates that allows the growth

of different fungal species leading into a distinct growth fingerprint).

111
Chapter 4: Commercial extract - characterisation

Figure 4.3. Kinetics of average well colour development (AWCD) and optical densities of (A)
bacterial (570 nm) and (B) fungal (490 nm) communities, respectively, from different compost
extracts and microbial preparations. 3MCE - three month old rumen compost extract, 9MCE -
nine month old rumen compost extract, 9MTCE - nine month old rumen compost extract plus
Nutri-Life 4/20TM inoculum, LS - Living soilTM, NL-4/20 - Nutri-Life 4/20TM (n = 3) sampled
after 24 h of incubation.

The average well colour for bacterial (Ecoplate) growth of the compost

extract amended with Nutri-Life 4/20TM inoculum was not different to that of

unamended compost extract, however, higher fungal (FF microplate) growth was

noted in nine month old rumen compost extract plus Nutri-Life 4/20TM inoculum

compared to nine month old rumen compost extract (Fig. 4.3). The added

bacterial species in nine month old rumen compost extract plus Nutri-Life 4/20TM

inoculum might have been outcompeted and/or eliminated by the microbiota

already present in the compost extract. However, higher fungal (FF microplate)

112
Chapter 4: Commercial extract - characterisation

growth was noted in nine month old rumen compost extract plus Nutri-Life

4/20TM inoculum compared to nine month old rumen compost extract (Fig. 4.3B).

Significant differences among treatments in average utilisation of specific

substrate guilds were evident when BiologTM (Ecoplate and FF microplate) data

were analysed by guild categories (Fig. 4.4). The Living soilTM culture displayed

the most even growth across all carbon substrates on the Ecoplate cultures. Of all

guild categories, no difference in bacterial substrate utilisation among nine month

old rumen compost extract, nine month old rumen compost extract plus Nutri-

Life 4/20TM inoculum and Nutri-Life 4/20TM was apparent except nine month old

rumen compost extract plus Nutri-Life 4/20TM inoculum averaged higher

utilisation of polymers than the other treatments (Fig. 4.4A). Similarly, the FF

microplate data of three month old rumen compost extract and Living soilTM

averaged higher utilisation of all guild categories of carbon substrates than the

other treatments (Fig. 4.4B). The fungal substrate utilisation of nine month old

rumen compost extract and nine month old rumen compost extract plus Nutri-

Life 4/20TM inoculum as shown by FF microplate data did not vary, however,

they averaged higher utilisation of amino acids, amines/amides and miscellaneous

substrates.

113
Chapter 4: Commercial extract - characterisation

Figure 4.4. Average carbon substrate utilisation from different substrates by samples from three
month old rumen compost extract (3MCE), nine month old rumen compost extract (9MCE), nine
month old rumen compost extract plus Nutri-Life 4/20TM inoculum (9MTCE), Living soilTM (LS)
and Nutri-Life 4/20TM (NL-4/20) consuming (A) 32 (Biolog Ecoplate) and (B) 96 (Biolog FF
microplate) different food sources. Samples were collected from 72 h incubations of either Biolog
plate and utilisation was measured as the average optical density across all substrates witihin each
guild which included 7 different carbohydrates, 9 carboxylic acids, 4 polymers, 6 amino acids, 2
amines/amides and 3 miscellaneous for Ecoplate while 44 different carbohydrates, 17 carboxylic
acids, 5 polymers, 13 amino acids, 6 amines/amides and 10 miscellaneous for FF microplate.

Biolog data represent the physiological profile of culturable

microorganisms (Garland & Mills, 1991; Widmer et al., 2001; Zak et al., 1994).

The PCA of Ecoplate data explained 66% (PC1) and 10% (PC2) of the total

variance, whereas the PCA of FF microplate data explained 89% (PC1) and 2%

(PC2) of the total variance (Fig. 4.5). Principal component analysis revealed two

114
Chapter 4: Commercial extract - characterisation

distinct clusters clearly for both bacteria and fungi (Fig. 4.5), with Living soilTM

and three month old rumen compost extract separated from nine month old rumen

compost extract, nine month old rumen compost extract plus Nutri-Life 4/20TM

inoculum and Nutri-Life 4/20TM. This suggests a large difference in catabolic

capacity between the microbial communities of these cultures (Fig. 4.5).

Relatedness of liquid cultures within either group indicates that the functional

groups of microbiota present in each culture are similar, possessing similar

potential for substrate utilisation. However, the overall shift in community

structure could not be attributed to the changes in the abundance of any particular

functional group in accord with Garland (1999), as indicated by the loadings of

the first principle component (data not shown).

115
Chapter 4: Commercial extract - characterisation

Figure 4.5. Principal component analysis (PCA) of optical densities produced by microbial
activities of five different cultures consuming 32 (Biolog Ecoplate) and 96 (Biolog FF microplate)
different food sources including the control; A and B representing bacterial and fungal diversity.
3MCE - three month old rumen compost extract, 9MCE - nine month old rumen compost extract,
9MTCE - nine month old rumen compost extract plus Nutri-Life 4/20TM inoculum, LS - Living
soilTM, NL-4/20 - Nutri-Life 4/20TM (n = 3), all of which sampled at 24 h after incubation.

Since the Living soilTM formulation was derived from a compost based on

cow manure (which was further vermicomposted), it was not surprising to find

similarity of its bacterial communities with those of three month old compost

extract, although the fungal communities between those two varied (Fig. 4.5).

While the fungal substrate utilisation pattern of Living soilTM clustered with that

of nine month old rumen compost extract, nine month old rumen compost extract

plus Nutri-Life 4/20TM inoculum and Nutri-Life 4/20TM, the fungal community of

116
Chapter 4: Commercial extract - characterisation

three month old compost extract was distinctly separated from those of other

treatments, suggesting a dissimilarity in species composition of microflora

between them (Fig. 4.5). Relatedness of nine month old rumen compost extract

and nine month old rumen compost extract plus Nutri-Life 4/20TM inoculum was

understandable, with little influence from the addition of the microbial

amendment to the compost extract. The functional diversity of Nutri-Life 4/20TM

was similar to that of nine month old compost extracts. This result was

unexpected, and no explanation is offerred.

3.4. Microscopic analyses

The methods used in the Soil Foodweb Institute (SFI) analysis are not

defined, but presumably are based around microscopic assessment of cell

numbers (from indications in SFI tutorial information). The SFI results suggest

that rumen based compost extract is bacterial dominated as compared to the

fungal dominated Living soilTM, with both extracts supplying only minimal

amounts of available N to the plants (Table 4.2).

Table 4.2. Sample analysis data of nine month old rumen compost extract (9MCE) and LS
(Living SoilTM) provided by Soil Foodweb Institute, NSW, Australia.
Expected range
Attributes 9MCE Comment LS Comment Low High
Sample size (mL) 1 - 1 - - -
Active bacteria (µg/mL) 179 above range 80.4 in range 10 150
Total bacteria (µg/mL) 11136 above range 17408 above range 150 3000
Active fungi (µg/mL) 123 above range 417 above range 2 10
Total fungi (µg/mL) 162 above range 427 above range 2 20
Hyphal diameter (µm) 4.5 - 4.5 - - -
Protozoa

Flagellates (numbers/g) 7 Low 0 Low 1000 -


Amoebae (numbers/g) 0 Low 0 Low 1000 -
Ciliates (numbers/g) 0 Low 0 Low 20 50
Total nematodes 0 Low 0 Low 2 10
Total fungi to total bacteria 0.01 Good 0.02 Good 0.01 0.1
Active to total fungi 0.76 High 0.98 High 0.1 0.25
Active to total bacteria 0.02 Low 0.005 Low 0.1 0.25
Active fungi to active bacteria 0.69 Low 5.18 High 0.9 1.1
Plant available N supply (lbs/ac) <5 - <5 - - -

117
Chapter 4: Commercial extract - characterisation

The SFI results, however, are not consistent with the results of the

bacterial and of the fungal plate counts (Table 4.1). This apparently pinpoints the

limitations of direct count method using a microscope (Scheuerell, 2004),

although it is possible that the samples had altered during shipment. Additionally,

Bloem et al. (1995) indicated that living and dead cells could not be differentiated

by direct count method, thereby leading to overestimation.

3.5. PCR and DGGE

A UPGMA dendrogram was generated to express the relatedness of

microbial communities as similarity clusters (Fig. 4.6). Replicate variability of

both bacteria and fungi was higher in nine month old compost extracts than that

of other cultures, the latter showing more than 90% similarity in their replicated

DNA (Fig. 4.6).

118
Chapter 4: Commercial extract - characterisation

Figure 4.6. UPGMA (Unweighted Pair Group Method with Arithmetic mean algorithm)
dendrogram for (A) bacteria and (B) fungi present in 3MCE, 9MCE, 9MTCE, LS and NL-4/20.
3MCE - three month old rumen compost extract, 9MCE - nine month old rumen compost extract,
9MTCE - nine month old rumen compost extract plus Nutri-Life 4/20TM inoculum, LS - Living
soilTM, NL-4/20 - Nutri-Life 4/20TM (n = 2), all of which sampled at 24 h after incubation.

A dissimilarity in bacterial community composition between nine month

old compost extract and that amended with Nutri-Life 4/20TM inoculum was

revealed, indicating that although overall activity and functional diversity were

not altered by the amendment, species composition was.

Curiously, the amended nine month old compost extract was more similar

to three month old rumen compost extract with approximately 85% homology,

than to the non-amended extract (Fig. 4.6A). The addition of Nutri-Life 4/20TM

inoculum presumably added bacterial species similar to those present in three

month old rumen compost extract and to a lesser extent to those present in Living

119
Chapter 4: Commercial extract - characterisation

soilTM. These species must have been absent or below limit of detection by the

DGGE method in the nine month old rumen compost extract.

The 16S rDNA and Internal Transcribed Spacer region based DGGE

analyses showed significantly greater bacterial community diversity of rumen

based compost extracts and Living soilTM compared to that of Nutri-Life 4/20TM;

while fungal communities were more similar in their diversity between treatments

(Table 4.3). The calculated values for Shannon diversity index (H‘) and

equitability index (J) obtained from the bacterial DGGE profile (Table 4.3) were

not different between the treatments of three month old rumen compost extract,

nine month old rumen compost extract, nine month old rumen compost extract

plus Nutri-Life 4/20TM inoculum and Living soilTM, but this group had

significantly higher values for diversity and evenness compared to those of Nutri-

Life 4/20TM. This suggests the lowest diversity of bacterial community in Nutri-

Life 4/20TM.

Table 4.3. Diversity and equitability of microbial communities in different liquid cultures
compared. Shannon-Weaver diversity index, H‘ (Shannon and Weaver, 1949) and equitability, J
(Pielou, 1966) and number of bands, S were generated from DGGE profiles of PCR amplified
16S rDNA and ITS genes from all samples collected at 24 h of incubation. 3MCE - three month
old rumen compost extract, 9MCE - nine month old rumen compost extract, 9MTCE - nine month
old rumen compost extract plus Nutri-Life 4/20TM inoculum, LS - Living soilTM, NL-4/20 - Nutri-
Life 4/20TM (n = 2), all of which sampled at 24 h after incubation. Within columns, means with
the same letter are not significantly different according to Paired T-test (p< 0.05).
Bacteria Fungi
Treatments H' J S H' J S
3MCE 2.95 ± 0.03b 0.82 ± 0.00b 36 ± 1.00c 2.93 ± 0.00a 0.87 ± 0.00a 29 ± 0.50b
9MCE 2.38 ± 0.29b 0.74 ± 0.06b 25 ± 3.00b 2.56 ± 0.28a 0.86 ± 0.05a 20 ± 3.00ab
9MTCE 2.81 ± 0.03b 0.80 ± 0.01b 33 ± 0.00bc 2.17 ± 0.36a 0.80 ± 0.05a 15 ± 4.00a
LS 2.83 ± 0.05b 0.80 ± 0.01b 34 ± 0.50c 2.07 ± 0.12a 0.74 ± 0.03a 17 ± 0.50ab
NL-4/20 1.09 ± 0.13a 0.39 ± 0.05a 16 ± 0.00a 2.33 ± 0.00a 0.81 ± 0.00a 18 ± 0.00ab

However, the significantly higher number of bands, as indicated by

species richness (S), in the DGGE gels of three month old rumen compost extract

and Living soilTM implies greater diversity of the bacterial communities of these

120
Chapter 4: Commercial extract - characterisation

two treatments in comparison to the remaining treatments (Table 4.3). Higher

values of S representing greater bacterial diversity in three compared to nine

month compost extract agreed with the results of Biolog data as discussed earlier.

Unlike bacterial community analysis, no differences in Shannon-Weaver

diversity and evenness were found between the treatments for fungi, however,

higher fungal species richness was evident in three month old rumen compost

extract compared to nine month old rumen compost extract plus Nutri-Life

4/20TM inoculum (Table 4.3). The addition of Nutri-Life 4/20TM inoculum, which

is claimed to be rich in beneficial fungi including Trichoderma, to nine month

compost extract failed to increase the species richness (S) of fungal species

compared to three month old rumen compost extract (Table 4.3). This result can

be explained as indicating either that the species added were already present in

the compost extract, or that the added species were outcompeted and eliminated

by the microbiota of the compost extract, as indicated by other measures of

microbial activity, and Biolog measure of diversity or they were not detected by

DGGE.

Unlike the production of a pure culture (e.g. of Rhizobium, for legume

inoculation; or of a fungal plant pathogen, for disease studies), the production of

microbially enhanced ‗soil health‘ products is an ‗inexact science‘ due to the

enormous possible microbial variation. It is difficult to offer criteria as to the

‗quality‘ of the product. Rather, there is an inherent assumption that the microbial

assemblage present in aerobic compost is beneficial for plant growth (e.g.

Ingham, 2005; Pant et al., 2009). However, this assemblage will differ with

compost condition (e.g. age) and incubation conditions (e.g. aeration, food

sources) (Carpenter, 2005). Commercial products, such as Nutri-Life 4/20TM and

121
Chapter 4: Commercial extract - characterisation

Living soilTM, seek to provide storable forms of a set of beneficial microbiota,

although changes in the assemblage during storage are likely. Further, incubation

conditions (e.g. anaerobic vs. anaerobic) are also likely to result in a change in

the microbial assemblage.

4. Conclusions

The Biolog and DGGE tools were useful in characterising the microbial

assemblages of the liquid cultures. The three month old rumen compost extract

and Living soilTM were closely related and functionally more active compared to

nine month old rumen compost extract, nine month old rumen compost extract

plus Nutri-Life 4/20TM inoculum and Nutri-Life 4/20TM which showed

relatedness to each other. The addition of Nutri-Life 4/20TM inoculum in nine

month old rumen compost extract plus Nutri-Life 4/20TM inoculum did not alter

the functional microbial activity compared to the original extract. Fungal

diversity did not vary, but bacterial diversity was higher in all cultures compared

to Nutri-Life 4/20TM. The potential role of such biofertilisers in soil functioning

and crop health needs to be explored.

122
Chapter 5: Compost extract and plant nutrition

Chapter 5

Microbially enhanced compost extract: does it

increase solubilisation of minerals and

mineralisation of organic matter and thus improve

plant nutrition?

A version of this chapter has been submitted to Crop and Pasture Science (currently under
review)

Abstract

Microbially enhanced compost extracts are increasingly being used to

improve soil and crop ‗health‘. The aim of this research was to assess potential

mechanisms by which compost extracts may influence plant growth. Tomato

(Lycopersicum esculentum cv. Tiny Tim) and sorghum (Sorghum bicolor cv.

Sweet Jumbo LPA) were used as test crops. The effect of rumen content compost

extracts (aerated and non-aerated, sterilised or not) on plant growth was

compared to inorganic fertiliser and a water-control in greenhouse and growth

chamber pot trials using various growing media. Soil type was the primary factor

in determining growth rate of crops. Compost extracts produced by either

extraction method had similar effects on plant growth, and there was no

difference between sterilised and non-sterilised treatments. Thus the noted growth

benefit of compost extracts was not directly biological in nature. The positive

123
Chapter 5: Compost extract and plant nutrition

impact of compost extract application on plant growth was ascribed to a

nutritional effect, attributed to the high doses of compost extract application used

(2 L pot-1, equivalent to 34,000 L ha-1), delivering 3.4 g of N pot-1.

1. Introduction

Compost extract, popularly known as ‗compost tea‘, is an extract of

compost incubated with or without a microbial food source with the aim of

extracting soluble nutrients and plant beneficial microbes into the solution (Diver,

2002). Although there is an extensive anecdotal evidence to support the claim of

improved plant growth in response to compost extract application, studies

conducted under controlled conditions are limited. For example, Egamberdiyeva

(2007) reported evidence of higher nutrient uptake and a growth stimulatory

effect of microbial inoculation (with growth promoting rhizobacteria, namely,

Pseudomonas alcaligenes, Bacillus polymyxa and Mycobacterium phlei) on

maize in nutrient deficient soils. However, the direct cause of this response was

not elucidated – the effect could be due amongst others to a direct fertilisation

effect, increased mineralisation/solubilisation, or a disease suppression effect.

For example, Ekin (2010) reported increased shoot biomass and seed

yield of sunflowers associated with a soil application of phosphate solubilising

bacteria (Bacillus M-13) and a phosphorus fertiliser, compared to fertiliser alone.

This result was ascribed to the effect of soil microbiota on soil mineral

solubilisation at a level significant to plant growth.

The extracted mineral nutrients can improve soil fertility directly

(Campbell, 2006; Merrill et al., 1997; Scheuerell & Mahaffee, 2002). Other

claims include a role for the extracted microbiota in improved mineralisation of

124
Chapter 5: Compost extract and plant nutrition

soil organic matter and solubilisation of soil minerals, chelation of ions (Janzen et

al., 1995), suppression/biocontrol of certain plant root and foliar diseases (Bernal-

Vicente et al., 2008; Haggag & Saber, 2007), and microbial production of plant

growth promoting hormones such as auxins (Garcia et al., 2002), or cytokinin-

like substances (Arthur et al., 2001).

However, many reports of the use of compost extract involve high rates of

application, such that a direct nutritional effect will dominate. For example,

Hargreaves et al. (2008) reported that compost extract (produced at the ratio of 1:

10 w/v) applied at weekly intervals to a raspberry crop at the rate of 150-300 mL

per plot (2.4 m2) per application (NOSB, 2004) promoted the growth of

raspberries as effectively as applying compost at the rate of 150 kg total N per ha

in 2004 and 75 kg N per ha in the following years to the soil. However, 27

applications made over 3 years will have delivered approximately 4000 - 8000
-2
mL 2.4m (i.e. equivalent to 17,000 - 34,000 L ha-1, derived from 1.7 – 3.4 t

compost), however, the actual supply of N from these compost extracts has not

been indicated. In another study, Hargreaves et al. (2009) produced compost

extracts by steeping compost for 72 h at the ratio of 1: 5 (w/v), with application at

the rate of 1 L and 2 L per plot (of 2 m2 containing six strawberry plants) per

week continuously for twelve weeks in 2006 and for five weeks in 2007 (i.e.

120,000 L ha-1, or 24 ton ha-1 of compost). Compost extracts may thus be used to

maintain soil nutrient levels by growers who are restricted from using inorganic

fertilisers by the norms of organic farming, with the advantage of easier

distribution (e.g. through irrigation systems) than is possible for solid compost.

Compost extracts, however, can vary in composition. There are limited

studies comparing methods of preparation of compost extracts (Pant et al., 2009).

125
Chapter 5: Compost extract and plant nutrition

For example, compost extracts can be produced following aerobic (aerated

compost extract) or anaerobic methods (non-aerated compost extract),

presumably resulting in distinctly different microbial populations, however there

is a lack of evidence as to the superiority of one or other method for agricultural

purposes (Brinton et al., 2004). Ingham (1999a) has advocated the use of aerobic

compost extracts, whereas other studies have found a significantly positive effect

of non-aerated compost extracts on plant growth due to disease suppression

(Scheuerell & Mahaffee, 2004; 2006; Tranker, 1992; Weltzien, 1991).

The present experiment was designed to test two hypotheses: (a) that

application of compost extract to the soil results in an enhancement of the

microbial population of the soil, and (b) that microorganisms present in compost

extract may enhance solubilisation of relatively inaccessible inorganic pools and

increase mineralisation of organic matter to release mineral nutrients, thereby

improving soil fertility. Two pot trials were conducted to determine the effects of

extraction methods (aerobic and non-aerobic) on the growth of tomato and

sorghum supplied with extracts (either sterilised or non-sterilised) derived from a

cattle rumen content compost. In one trial, the soil medium contained sand and

leached compost (to test mineralisation capacity), while in the second trial two

soils were used, both of low nutrient availability and organic matter content (to

test solubilisation capacity), with a rotation crop planted into the root residues of

the first crop (to test mineralisation capacity).

Tomato was chosen as a likely horticultural crop for ‗compost tea‘

application, being Australia‘s second largest commercial vegetable crop, with

Queensland being the main producer of fresh market tomatoes (DPIF, 2000).

126
Chapter 5: Compost extract and plant nutrition

2. Materials and methods

2.1. Compost and compost extract

Cattle rumen content (paunch material) was used for compost given the

relative uniformity of this material (e.g. relative to urban green waste) and its

local abundance (with up to 1700 head of cattle slaughtered per day in

Rockhampton abattoirs). A commercial composting operation in Rockhampton

utilises a thermophilic windrow composting process, with the windrows (2 m

height, 2.5 m width and 50 m length of material) turned mechanically at monthly

intervals, for nine months to ensure homogeneity and aeration. Four

consignments of nine month old compost were used in the course of this study.

Compost extracts were produced at the ratio of 1:10 (w/v) of compost

submerged in dechlorinated tap water following aerated and non-aerated methods

of extraction (at between 18˚C and 25˚C). Dissolved oxygen, measured with a

TPS WP 82 Dissolved Oxygen Meter (EnviroEquip, Australia), of aerated and

non-aerated compost extracts at 24 h of the incubation period was 71 % and 1.5

% of air-saturation, respectively. Addition of 0.5 % v/v of molasses and 1 % v/v

of soluble ‗fish and kelp hydrolysate‘ (Fish and Kelp Product, Australia) was

made prior to application of the compost extract to soil/growing media. Compost

extracts were prepared on the day of soil application.

Nutrient analysis of composite samples of the aerated and the non-aerated

compost extracts, i.e. representatives of all extracts used, was undertaken by the

Soil and Plant Laboratory, Wesfarmers CSBP Limited, Australia. The samples

were shipped in cooled liquid form (i.e. with ice blocks).

127
Chapter 5: Compost extract and plant nutrition

2.2. Experiment 1

The hypothesis was tested that non-sterile compost extract would enhance

plant growth over an equivalent amount of sterile compost extract through the

enhancement of soil mineralisation in an ‗artificial‘ soil of organic matter and

sand mix.

The growing media was prepared by mixing the thoroughly washed (i.e.

leached) sterilised rumen compost with sterilised sand at the ratio of 1: 5 v/v.

Materials (sand, compost, compost extract) were sterilised by autoclaving at

121˚C for 15 minutes. Three-week old tomato seedlings (Lycopersicum

esculentum cv. Tiny Tim) were transplanted one each into 13 cm diameter pots

and maintained in a growth chamber. Each pot contained 1 kg of the growing

media.

A completely randomised experimental design was implemented, with

five treatments: (i) control (i.e., equivalent amount of water only), (ii) aerated

compost extract (ACE), (iii) non-aerated compost extract (NCE), (iv) sterilised

aerated compost extract (S ACE) and (v) sterilised non-aerated compost extract

(S NCE). Pots were placed on a 45 cm by 25 cm grid, with eight replicate pots

per treatment. Treatments, including the water control, were applied twice at the

rate of 100 mL kg-1 compost-sand mix, the first being applied one week after

transplanting and the second a week later.

Two replicate trials (I and II) were conducted. In trial I, the growth

chamber was set at 28˚C, 720 µmol m-2 s-1 PAR and 75 % relative humidity

(RH). In trial II, the growth chamber was set with lower temperature and light

intensity, but with the same relative humidity as in trial I (24˚C, 480 µmol m-2 s-1

PAR and 75 % RH).

128
Chapter 5: Compost extract and plant nutrition

The growth parameters of plant height, stem diameter, number of leaves

and leaf chlorophyll concentration were measured at 10 day intervals starting

from 18 and 31 days after transplanting in trials I and II, respectively. Dry weight

of all plants (roots and shoot biomass) was measured following harvest at 45 and

51 days after transplanting in trials I and II, respectively.

2.3. Experiment 2

The hypothesis was tested that microbial enhancement of the soil could

serve to improve solubilisation of nutrients (and mineralisation of any soil

organic matter) and so improve plant growth.

A vertisol was sourced (courtesy of Eric Coleman, QDPIF) from a

Gracemere site (latitude: 23.44˚and longitude: 150.46˚) that had been maintained

as uncultivated grassland for over thirty years. A ferrosol was sourced (courtesy

of farmer Tony Wolfenden) from an organic sweet potato farm at Rossmoya

(latitude: 23.05˚ and longitude: 150.48˚), from a site that had been cultivated

regularly over the past decade. At both sites, soil was collected to a depth of 30

cm, and was transported as a single, homogenised sample (soil mixed using a

small front end loader). Prior to starting the experiment, replicated samples from

both soils were collected, homogenised individually and combined to obtain

representative samples and sent for nutrient analysis (Wesfarmers CSBP Limited,

Soil and Plant Laboratory, Australia).

Soil was air dried and each pot (30 cm diameter, 40 cm height and 20 L

volume) was filled with 20 kg of vertisol or 22 kg of ferrosol. All pots were

weighed and watered to achieve field capacity (40 % v/v and 38 % v/v for the

vertisol and ferrosol, respectively). Pots were maintained in a greenhouse

129
Chapter 5: Compost extract and plant nutrition

(ambient temperature ranging from 18 to 40˚C with 60 to 100 % relative

humidity). Pots were spaced at 75 x 50 cm. Tomato (cv. Tiny Tim) seeds were

germinated and seedlings transplanted to achieve one plant per pot.

The experiment used a factorial design within a randomised complete

block design, with four application treatments applied to the two soil types, and

four replicate pots per treatment. Treatments were (i) aerated compost extract, (ii)

non-aerated compost extract, (iii) chemical fertiliser and, (iv) a water control. All

treatments were applied as a soil drench, with 0.5 L solution applied weekly to

each pot for four weeks (equivalent to approximately 8,500 L ha-1 per

application), beginning from one week after transplanting. The chemical fertiliser

treatment utilised the same volume of solution containing 0.9 g of ―Thrive‖

(Yates, NSW Australia), which consisted of 15: 4: 26 NPK and all other essential

nutrients. Culturable heterotrophic bacterial and fungal counts (cfu Log 10 mL-1)

of representative samples of aerated and non-aerated compost extracts (n = 3),

sampled immediately prior to the application to the pots, were determined

following the pour plate method (Harrower, 2006).

Tomato plants (the same cultivar as used in Experiment 1) were grown for

76 days, being watered to field capacity at weekly intervals. Soil respiration and

soil temperature were measured per pot (n = 3) 76 days after transplanting, to

estimate plant root and microbial activity in the soil using an EGM3 respirometer

(PP Systems International Ltd., Hertfordshire, UK). Plant height, stem diameter,

leaf number, leaf chlorophyll concentration (SPAD-502 meter, Minolta

Corporation, Japan), plant nutrient status, fruit yield and above ground biomass

were assessed at harvest on all plants (76 days after transplanting).

130
Chapter 5: Compost extract and plant nutrition

The microbial population of the soil was quantified in triplicate at 36 days

after tomato transplanting, that is, one day after the last compost extract

application, using a plate count method (Table 5.4) (Harrower, 2006). Soil was

sampled with a 2 cm diameter core sampler from the top 10 cm depth,

composited, and 1 g of representative soil sample was taken from each treatment.

A nitrogen budget was constructed. Shoot material, compost extract and

fertiliser, dried at 65˚C, were ground using a CQUniversity manufactured ball

mill grinder. Subsamples (10 mg) were packed in tin capsules. Sample nitrogen

elemental and isotopic content (natural abundance) was determined using an

automated continuous flow isotope cube utilising Dumas flash combustion

(Elementar, Germany) coupled with a continuous flow mass spectrometer

(Isoprime, England) at the University of Melbourne. Machine precision and

standardisation were established with acetanilide (Merck, Germany) as a tertiary

standard (10.36 %N and 0.2 ‰ δ15N) with analytical precision (1σ) of < 0.1 ‰.

2.4. Experiment 3

The hypothesis was tested that microbial enhancement of the soil could

serve to improve mineralisation of residual tomato roots, and hence improve plant

growth.

Following harvest of the tomato crop (removal of shoot system only, with

the soil left undisturbed) and a fallow period of one month, 10 sorghum seeds

(Sorghum bicolor cv. Sweet Jumbo LPA) were directly sown in each pot, and

thinned out after one week to maintain six plants per pot. From the previous

tomato trial, there was no significant effect of blocks on yield (above-ground

biomass). In that trial, blocks were arranged down the length of the greenhouse.

131
Chapter 5: Compost extract and plant nutrition

Blocks were rearranged across the width of the greenhouse to test for an

environmental gradient in this direction (i.e. a possible wind speed gradient away

from the windsock located at one edge of the greenhouse).

Two treatments, namely, aerated compost extract and sterilised aerated

compost extract, were applied at three week intervals at the rate of 5 mL per kg of

soil. The rate of compost extract application was reduced to minimise nutrient

input (equivalent to approximately 1700 L ha-1 per application). These treatments

were applied to two pots each of the original four replicates of the first (tomato)

rotation (i.e. across the two soils (vertisol and ferrosol) and four previous

treatments (control, aerated compost extract, non-aerated compost extract, and

chemical). The pots were randomly assigned treatments within three blocks.

Percent light interception at the base of the sorghum plant per pot was

measured using a Ceptometer, LAI-2000 (Decagon, USA), and leaf chlorophyll

concentration of the top third leaf per plant per pot was indexed using a SPAD-

502 meter (Minolta Corporation, Japan). Soil respiration (root + microbial

respiration) per pot (n = 3) was measured with an EGM3 respirometer just before

harvest. The vegetative growth (shoot biomass), as measured by above-ground

plant dry weight, of the forage sorghum crop was measured on all pots at 50 days

after sowing.

2.5. Statistical analyses

Data were analysed using the GLM procedure of GENSTAT (2008) and

statistical significance was evaluated at p< 0.05 and p< 0.001. The effect of

treatments on different parameters including plant biomass was analysed using

either one-way or two-way analysis of variance (ANOVA). For the glasshouse

132
Chapter 5: Compost extract and plant nutrition

trial, tomato shoot biomass (plant dry weight in g) was used as a covariance-

factor on the sorghum shoot biomass for the analysis of variance.

3. Results

3.1. Characterisation of compost extract

The soluble nutrient contents of compost extracts produced under aerated

and non-aerated conditions were similar (Table 5.1).

Table 5. 1. Nutrient (macro and micro) concentrations for aerated and non-aerated compost
extract samples analysed by Wesfarmers CSBP Limited, Australia. ‗Thrive‘ nutrients based on the
chemical composition supplied by manufacturer Yates and Co. Limited, Australia.

Nutrients (mg L-1) ACE NCE Thrive


NO3-N 0.80 1.10 157
+
NH4 -N 80.6 70.2 64.0
N as Urea - - 46.0
Total soluble N 81.4 71.3 267
P 186 182 71.0
K 334 334 463
Ca 146 166 -
Mg 29.1 26.5 9.00
S 95.5 101 -
Cu 0.34 0.15 0.09
Mn 0.42 0.54 0.71
Zn 1.28 1.86 0.36
Fe 8.24 9.36 3.20
B 0.24 0.24 0.09
Na 411 418 -
Cl 374 375 -
Mo - - 0.04
ACE - Aerated compost extract, NCE - Non-aerated compost
extract, Thrive - water soluble nutrient (chemical fertiliser).

In each compost extract, 3 kg f wt of compost at 60 % moisture (i.e. 1.2

kg d wt) was added to 30 L of water. Given a N content of the compost of 1.9 %

d wt, 22.8 g N was added to the 30 L extract in the form of compost. Each

compost extract also contained 1 % of Searles Fish and Kelp. Given a N content

of 10.3 % w/v in this product (as reported on label), 30.9 g N was added to the 30

L brew in the form of this product. Thus, a 30 L compost extract is expected to

133
Chapter 5: Compost extract and plant nutrition

contain 53.7 g N. In supplying the extract at the rate of 2 L per pot, 3.58 g N

should have been delivered.

3.2. Experiment 1: Tomato growth in sand-compost mix

3.2.1. Plant biomass

In both trials, total plant biomass (dry weight, shoot + root) was

significantly different (p< 0.001) between imposed treatments and the control

(Fig. 5.1A), however there was a non-significant difference among the compost

extract treatments (both sterilised and non-sterilised).

In all treatments receiving compost extracts, tomato plant shoot to root

ratio was increased compared to the control, but no difference was found between

compost extract treatments (Fig. 5.1B).

134
Chapter 5: Compost extract and plant nutrition

Figure 5. 1. Effect of compost extracts on (A) total plant biomass (g d wt) and (B) shoot: root
biomass of tomato plants in Experiment 1 grown in a thoroughly washed compost-sand mix at 1:
5 ratio in two replicate trials, harvested at 48 and 51 days after transplanting, respectively. Cont -
control, ACE - aerated compost extract, NCE - non-aerated compost extract, S ACE - sterilised
aerated compost extract and S NCE - sterilised non-aerated compost extract. Within a trial, means
with different letters indicate significant difference (p< 0.05). Vertical bars represent the standard
error of the mean (n = 8).

3.2.2. Tomato leaf chlorophyll concentration

In trial I (Fig. 5.2A), chlorophyll concentrations (SPAD Units) were

significantly higher in compost extract treated plants than in the control at 18, 28

(except for aerated compost extract) and 38 days after transplanting (p< 0.001, p<

0.01 and p< 0.05, respectively). In trial II (Fig. 5.2B), higher chlorophyll

concentrations were found in leaves of plants treated with both sterilised and non-

sterilised compost extracts (aerated and non-aerated) compared to the control at

31 (p< 0.001) and 41 (p< 0.05) days after transplanting.

135
Chapter 5: Compost extract and plant nutrition

However, in trial II, no difference in leaf chlorophyll concentrations

between treatments was recorded at 51 days after transplanting.

Figure 5. 2. Effect of compost extracts in Experiment 1 on chlorophyll concentration in leaves in


(A) trial I, and (B) trial II. Cont - non-fertilised, ACE - aerated compost extract, NCE - non-
aerated compost extract, S ACE - sterilised aerated compost extract and S NCE - sterilised non-
aerated compost extract. Means with different letters within a sampling date indicate significant
difference (p< 0.05) and vertical bars represent the standard error of the mean (n = 8).

3.2.3. δ15N of fertilisers (compost and compost extracts) and plants

grown in compost-sand mix

Mean values for δ15N content of the compost (at 15 ‰), used in the

compost-sand mix, were lower than that for the two compost extracts (at 17 ‰), a

result consistent with denitrification and volatilisation in the compost extracts,

however this difference was not significantly different due to replicate variability

136
Chapter 5: Compost extract and plant nutrition

(typical standard error of approx 0.5 ‰) (Fig. 5.3). However, plant δ15N was

similar in all three treatments, and significantly lower than that of the compost

extract.

Figure 5. 3. δ15N values of fertilisers applied including the rumen compost and of the tomato
plants grown in compost-sand mix (1: 5) under growth chamber conditions (Experiment 1). Cont -
δ15N of non-fertilised tomato plants and rumen compost, NCE - non-aerated compost extract, and
S ACE - sterilised aerated compost extract. Means with different letters indicate significant
difference (p< 0.0001) and vertical bars represent the standard error of the mean (n = 3).

3.3. Experiment 2: Tomato growth on soil

3.3.1. Soil

The two soils were characterised with respect to their basic properties

(Table 5.2). The ferrosol was comparatively nutrient poor, in terms of P and K,

and slightly more acidic than the vertisol.

137
Chapter 5: Compost extract and plant nutrition

Table 5. 2. Characterisation of vertisol and ferrosol used in the greenhouse trial.

Soil characterisitcs Vertisol Ferrosol


Total N (%) 0.15 0.22
NO3-N (mg kg-1) 9.00 7.00
+
NH4 -N (mg kg-1) 3.00 3.00
Total P (mg kg-1) 138 29.0
P Olsen (mg kg-1) 48.9 7.20
K (mg kg-1) 506 256
Organic C (%) 1.96 2.52
Fe (mg kg-1) 2613 2285
S (mg kg-1) 3.60 10.7
pH (H2O) 7.40 6.40
pH (CaCl2) 6.10 5.40
EC (dS m-1) 0.06 0.06
BD (kg L-1) 1.10 1.20
Texture 3.00 3.50
Gravel Nil Nil
Colour Dark brown Orange
Classification Heavy clay Clay
Soil texture was determined by the "Feel method" (CSBP lab)

3.3.2. Tomato growth, yield and chlorophyll concentration

At harvest (i.e. 76 days after transplanting), shoot biomass (g d wt) was

significantly lower (p< 0.001) for plants grown in the ferrosol than those grown

in the vertisol, irrespective of the amendment treatments (Fig. 5.4A). In both soil

types, shoot biomass was significantly lower in the control treatment (p< 0.001)

compared to the other three treatments, however there was no significant

difference among the three (two compost extract and one inorganic) treatments in

both vertisol and ferrosol soils (Fig. 5.4A). Although soil type showed a

significant effect (p< 0.001) on fruit yield per plant, there was no significant

difference in fruit yield among any of the treatments (Fig. 5.4B). There was a

proportionally greater benefit of non-aerated compost extract on leaf chlorophyll

concentration on the ferrosol than the vertisol (Fig. 5.4C). Chlorophyll

concentration was increased in plants treated with either compost extract, relative

to the control and the chemically treated plants (Fig. 5.4C).

138
Chapter 5: Compost extract and plant nutrition

Figure 5. 4. Data from Experiment 2 on (A) total above-ground biomass (g d wt per plant) of
tomato subject to four treatments after four months of planting on two soil types, (B) fruit yield of
tomato (g d wt per plant), (C) leaf chlorophyll concentration (SPAD unit) measured at 76 days
after transplanting of tomato. (i) Cont - non-fertilised, (ii) ACE - aerated compost extract, (iii)
NCE - non-aerated compost extract and (iv) Chem - chemical fertiliser. Within a given soil type,
means with different letters indicate significant difference (p< 0.05) and vertical bars represent the
standard error of the mean (n = 6).

3.3.3. δ15N of fertilisers and plants grown on soil

The δ15N isotopic composition of the inorganic fertiliser used was low, at

5 ‰, while that of the compost extract was high, at 18 ‰ (Fig. 5.5). The latter

139
Chapter 5: Compost extract and plant nutrition

result is consistent with denitrification occurring in the composting and extraction

processes, with 14N preferentially lost.

Figure 5. 5. δ15N values of tomato plants grown in ferrosol conducted in a greenhouse


(Experiment 2). Cont - non-fertilised, Chem - chemical fertiliser, ACE - aerated compost extract.
Means with different letters indicate significant difference (p< 0.0001) and vertical bars represent
the standard error of the mean (n = 3).

Tomato plants grown on a ferrosol without amendment possessed a δ15N

isotopic composition of approximately 12 ‰, while that of plants grown on the

same substrate but with addition of inorganic fertiliser was 6 ‰, and that of the

fertiliser itself was 5 ‰. This result indicates that (12-6)/ (12-5)*100 = 86 % of N

sourced by the plant was from the chemical fertiliser, with the remainder sourced

from the soil. The observed low contribution of N from the soil is consistent with

a low rate of N mineralisation and thus availability.

In contrast, the δ15N isotopic composition of plants treated with compost

extract (but no inorganic fertiliser) was 15 ‰, with that of the compost extract

itself, 18 ‰, indicating that (15-12)/ (18-12)*100 = 50 % of N sourced by the

plants in this treatment was sourced from the extract, although the δ15N values of

aerated compost extract treated plants was not significantly different to that of the

140
Chapter 5: Compost extract and plant nutrition

control plants (Fig. 5.5). Note that these calculations assume that the compost

extract was homogenous in its δ15N composition (e.g. soluble and insoluble

pools).

3.3.4. Microbial enumeration of aerated and non-aerated compost

extracts

The bacterial colony forming units (cfu) load of non-aerated compost

extract was approximately two orders of magnitude lower than that of aerated

compost extract, at all assessment dates, while the fungal load was similar in the

two extract preparations (Table 5.3). The addition of 2 L of aerated compost

extract per pot will have involved the addition of approximately 2 x 1012 cfu of

bacteria and approximately 7 x 107 cfu of fungi to each pot.

Table 5. 3. Enumeration of bacteria and fungi populations in representative samples of ACE and
NCE. The compost extracts were produced at the ratio of 1:10 w/v and amended with 0.5 % v/v of
molasses and 1 % v/v of soluble ‗fish and kelp hydrolysate‘, and were sampled immediately prior
to the application to the pots. Results are expressed as cfu Log 10 mL-1.

Application of compost extracts


CEs Population Week 1 Week 2 Week 3 Week 4
ACE Bacterial (cfu Log 10 mL-1) 7.24 ± 0.03a 9.47 ± 0.01b 9.46 ± 0.01b 10.40 ± 0.01c
Fungal (cfu Log 10 mL-1) 4.40 ± 0.10a 4.52 ± 0.04a 4.20 ± 0.10a 6.10 ± 0.10b
NCE Bacterial (cfu Log 10 mL-1) 5.81 ± 0.13a 9.13 ± 0.05c 8.83 ± 0.05c 8.46 ± 0.04b
Fungal (cfu Log 10 mL-1) 4.16 ± 0.16a 4.20 ± 0.10a 6.36 ± 0.23b 6.00 ± 0.00b
± values are SE of the means, n = 3 and means with different letters within rows are significantly
different at p< 0.05 in compost extracts (ACE and NCE)

3.3.5. Soil microbial populations

The population (log values) of culturable bacteria was significantly higher

in compost extract (both aerated compost extract and non-aerated compost

extract) treated soils, 36 days after the transplanting of tomato, compared to

control and chemical fertiliser treatments (p< 0.001) (Table 5.4). Consistent with

the fungal count in the compost extract itself, there were no significant

141
Chapter 5: Compost extract and plant nutrition

differences in either soil in terms of fungal populations in the two compost

extract treatments. Additionally, there was no significant difference in fungal load

between any of the four treatments including control and inorganic chemical

fertilisation (Table 5.4).

Table 5. 4. Effect of compost extracts on microbial (bacterial and fungal) populations (cfu Log 10
mL-1) in two soil types sampled at 36 days after transplanting (a day after the last compost extract
application) from the tomato experiment conducted in the greenhouse (Experiment 2). Means
with different letters within rows are significantly different at p< 0.05 (n = 3) in different
treatments, namely, Cont - control, ACE - aerated compost extract, NCE - non-aerated compost
extract, and Chem - chemical fertiliser. Results are expressed as cfu Log 10 mL-1.

Treatments
Soils Microorganisms Cont ACE NCE Chem
Vertisol Bacterial (cfu Log 10 mL-1) 4.67 ± 0.03a 6.27 ± 0.01c 6.19 ± 0.01c 4.88 ± 0.04b
Fungal (cfu Log 10 mL-1) 4.10 ± 0.10a 4.26 ± 0.14a 4.20 ± 0.10a 4.16 ± 0.16a
Ferrosol Bacterial (cfu Log 10 mL-1) 5.28 ± 0.03b 6.25 ± 0.01c 6.20 ± 0.01c 5.16 ± 0.01a
Fungal (cfu Log 10 mL-1) 4.10 ± 0.10a 4.20 ± 0.10a 4.36 ± 0.18a 4.10 ± 0.10a
± values are SE of the means, n = 3 and means with different letters within rows are significantly different
at p< 0.05 in various treatments, Cont - control, ACE - aerated compost extract, NCE - non-aerated
compost extract, Chem - chemical fertiliser.

3.3.6. Soil respiration and temperature

The measurement of soil respiration at harvest (76 days after

transplanting) represents a sum of both soil microbial and root respiration. There

was no significant difference evident in respiration rate, either between soil types

or treatments (data not shown).

Soil temperature (monitored 30 cm below the soil surface) did not differ

significantly between any of the treatments and soil types (data not shown).

3.3.7. Mass balance of nutrients

The four applications of 0.5 L of aerated compost extract (total of 2 L per

pot) resulted in an application of approximately 162 mg soluble N as NH4+ and

142
Chapter 5: Compost extract and plant nutrition

NO3−, 371 mg soluble P and 668 mg soluble K, respectively to each pot. An

additional 3580 mg insoluble N was present in the compost extract.

A mass balance of N, as an example, for each treatment in the vertisol was

calculated and is presented in Table 5.5. The aerated compost extract treatment,

for which there was an above-ground biomass of 130 g d wt, and given an

estimated root to shoot ratio of 1 and N content of 1.1 %, the calculated plant N

content was 2860 mg (Table 5.5). Thus, the N provided in the compost extract

was equivalent to approximately 125 % (N) of that in the plant.

Table 5. 5. Mass balance of N in compost extract and tomato crop grown in vertisol.

N added per pot Cont ACE NCE Chem


Soluble N added (mg) 0.00 162 143 534
Insoluble N added (mg) 0.00 3418 3437 0.00
Total N added (mg) 0.00 3580 3580 534
N in Plants
Aboveground biomass(d wt in g) 104 130 135 133
Assumed root: shoot ratio 1.00 1.00 1.00 1.00
% N in plants 1.00 1.10 1.70 1.10
Calculated plant N (mg)* 2080 2860 4590 2926
Added N/plant N X 100 (%) 0 125 78 18
*calculated as aboveground biomass X 2 X % N, Cont - control, ACE - aerated compost extract,
NCE - non-aerated compost extract, Chem - chemical fertiliser.

3.4. Experiment 3: Sorghum crop rotation

3.4.1. Plant growth, shoot biomass, chlorophyll concentration and light

interception

There were no visual differences between the growth of plants treated

with aerated and sterilised aerated compost extracts in either soil type while a

dramatic difference (p< 0.001) was observed in above-ground sorghum biomass

(harvested 50 days after sowing) between soil types (Fig. 5.6A).

143
Chapter 5: Compost extract and plant nutrition

Figure 5. 6. Data on sorghum from Experiment 3 on (A) above-ground dry weight biomass per
plant, (B) chlorophyll concentrations of leaves and (C) percent light interception (45 days after
sowing) grown in vertisol or ferrosol, according to treatments applied to the sorghum crop (x axis)
and to treatments applied to the previous rotation (tomato crop). Cont - non-fertilised, ACE -
aerated compost extract, NCE - non-aerated compost extract and Chem - chemical fertiliser.
Means with different letters across soil type and previous treatments indicate significant
difference (p< 0.05) and vertical bars represent the standard error of the mean (n = 6).

144
Chapter 5: Compost extract and plant nutrition

However, above-ground biomass did not differ between treatments

imposed on the sorghum plants (Fig. 5.6A), or between the treatments imposed

on the preceding tomato crop as shown by the covariance test (p = 0.98). For

percent light interception (Fig. 5.6C), significant difference was again found only

between soil types (p< 0.001).

There was no significant difference in sorghum leaf chlorophyll

concentration between any treatments (Fig. 5.6B).

4. Discussion

4.1. Effects of treatments on growth and nutrition of tomato plants

In the tomato growth trial (Experiment 1), conducted in sand-compost

mix, the lack of a plant growth difference between the sterilised and non-

sterilised treatments is consistent with the interpretation of a direct nutritional

benefit from the applied compost extract, rather than from a secondary benefit of

the microbial activity, such as mineralisation of organic matter in the sand-

compost mix.

The mass balance of N (Table 5.5) is again consistent with the

interpretation that the differences in plant growth between the applied treatments

is due to the nutrients supplied in the compost extract itself, rather than an

enhancement of soil mineralisation by the enhanced soil microbiota (Experiment

2).

The soluble and insoluble nutrient concentration of compost extracts

produced under aerated and non-aerated conditions was similar (Table 5.1, 5.5).

However, the N present in insoluble components of the compost extract (e.g.

microbial biomass) was approximately 20 times higher than in the soluble

145
Chapter 5: Compost extract and plant nutrition

fraction (Table 5.5). Potentially the nutrients in the microbial biomass would not

be immediately available for plant growth, and a greater plant response might be

expected from autoclaved extracts. As there was no such difference, the majority

of the applied microbial biomass presumably dies on application to the compost-

sand mix, freeing nutrients for plant growth.

No differences between aerated and non-aerated compost extract

treatments were observed in the current study in terms of growth (e.g. d wt

biomass) of tomato plants (Fig. 5.1 and Fig. 5.4). Pant et al. (2009) tested the

growth of pak choi under greenhouse conditions treated with aerated, non-aerated

and microbially enhanced vermicompost extracts prepared at the ratio of 1: 10

(w/v) and applied at the rate of 150 mL pot-1 for four continuous weeks,

compared to an aerated water-control. They also observed non-significant

differences in growth for plants treated with aerated and non-aerated compost

extracts, with all vermicompost regimes increasing above-ground fresh weight of

pak choi compared to the water-control.

In Experiment 1, tomato plants receiving compost extracts had a higher

shoot to root ratio compared to the control plants (Fig. 5.1B). An increased shoot:

root ratio is consistent with a relief of a nutrient limitation in the extract treated

plants (e.g. Ericsson, 1995).

The higher shoot to root ratio of tomato plants in the first trial compared

to the second trial could be attributed to the lower light level employed in the

second experiment.

146
Chapter 5: Compost extract and plant nutrition

4.2. Effects of treatments on tomato leaf chlorophyll concentration

In both trials I and II of Experiment 1, chlorophyll concentrations were

significantly higher in both sterilised and non-sterilised compost extract treated

plants compared to control. Fayed (2010a) also observed higher leaf chlorophyll

concentration in pomegranate trees treated with two foliar applications of

compost extract mixed with antioxidants at the rate of five litres per tree. Such

observations are consistent with a N or Fe fertilisation effect by the compost

extract.

However, leaf chlorophyll concentrations between treatments did not

differ 51 days after transplanting in trial II. The gradual increase in chlorophyll

concentration in plants of the control treatment must reflect the mineralisation of

organic matter present in the soil mix.

Higher leaf chlorophyll concentration was found in the first trial

compared to that in the second trial (Fig. 5.2B), a result which could be due to

higher light intensity and higher temperature employed in the first trial. Low

temperature and low light intensity can lead to an inhibition or impairment of

formation of chlorophyll in tomato (Allen & Ort, 2001; Hu et al., 2006).

Plants supplied with compost extract had a significantly lower δ15N level

than that of the compost extracts itself, suggesting that plants sourced N from the

mineralisation of the solid rumen compost in addition to the liquid compost

extracts (Fig. 5.3). Presumably the N of the compost extracts remained

unavailable to the plant as it was within microbial biomass.

In Experiment 2, the effect of compost extract on leaf chlorophyll

concentration (measured 76 days after transplanting) varied with soil type, with a

significant interaction (p< 0.001) between soil type and treatments noted (Fig.

147
Chapter 5: Compost extract and plant nutrition

5.4C). The significant interaction effect was attributed to the lack of response of

tomato (in terms of chlorophyll concentration) to the chemical fertiliser treatment

on the ferrosol. Chlorophyll concentration in tomato plants grown in the vertisol

was higher compared to those grown in the ferrosol (Fig. 5.4C), which could be

related to the higher nutrient availability in the former, especially, phosphorus

and potassium (Table 5.2).

4.3. Influence of soil types on crop growth

Soil type had a stronger influence on plant growth than the influence of

soil amendments used in the present study. In Experiment 2, vertisol produced

higher shoot biomass of tomato plants compared to those grown in the ferrosol,

which was consistent with the higher fertility (especially P) of the former soil.

Chemical fertiliser supported increased growth on ferrosol, however only 36 mg

P was given while the difference in soil P in the pots was 2200 g.

The grazing soil from which vertisol was collected has no known history

of fertiliser addition and is likely to be deficient in both Zn and B. Since compost

extracts have concentrations of Zn and B that are substantially higher than those

of the commercial fertiliser, plants grown in this soil are highly likely to respond

to the addition of the trace elements present in the extracts.

The vertisol also supported greater sorghum plant growth (above-ground

dry weight biomass) and high light interception compared to the ferrosol (Fig.

5.6A, 5.6C). This result is again attributed to the significantly greater nutritional

status of the vertisol in comparison to the ferrosol (Table 5.2).

148
Chapter 5: Compost extract and plant nutrition

4.4. Nutritional and microbial supply from compost extracts

The measured dissolved N (as ammonium and nitrate) content of the

extract represented only a small fraction (5 % and 4 % for aerated and non-

aerated compost extract, respectively) of the total N content (Table 5.1). The

difference (e.g. 3418 N mg per 2 L of aerated compost extract) must either be

present as microbial biomass, or insoluble or soluble but organic forms of N in

the compost extract, and/or this N was lost from the compost extract through
15
denitrification and volatilisation. The observed microbial load and the high N

values of the compost extract are consistent with both processes occurring.

Given the rate of extract application (2 L pot-1), the level of plant nutrients

in the compost extracts was certainly significant in terms of plant nutrition. For

example, the harvested (above-ground) biomass in Experiment 2 was

approximately 100 g d wt per pot. The average N concentration of above-ground

biomass across all treatments was 1.2 % (from elemental analyzer – mass

spectrometry analysis, data not shown). Given this, harvested plant biomass

contained 1.2 g N (per pot). Assuming a conservative shoot root ratio of 1: 1, the

total plant N content was 2.4 g per pot. Thus, a direct fertilisation effect is

expected for the extract treatments (3.4 g N delivered per pot).

The chemical fertiliser treatment was intended as a control exercise on the

effect of nutrient addition, as opposed to other possible effects of extract addition

(e.g. mineralisation). The chemical fertiliser application involved a similar order

of magnitude of soil inorganic elements per pot as the compost extract treatments,

except for N (which at 534 mg N pot-1 was much higher than the soluble

inorganic N pool in the compost extracts (Table 5.1), but also much lower than

149
Chapter 5: Compost extract and plant nutrition

the total N pool of the extracts). Little difference existed in the level of assessed

soluble nutrients between aerated and non- aerated compost extracts (Table 5.1).

However, the four compost extracts, prepared at weekly intervals, were

not consistent in their bacterial or fungal loads, with differences of up to two

orders of magnitude between batches. Of course, some variation is expected,

given the non-sterile, non-controlled incubation conditions and variation in the

compost inoculum, however this level of variation is large. Further, such

variation in bulk bacterial and fungal numbers also hints at variation in species

composition as well. More attention to the incubation process is recommended to

reduce the observed variation in final microbial load.

4.5 Effects of treatments on soil microbial properties

In Experiment 2, although there was a difference in bacterial counts in

compost extract treated soils compared to the control and chemical treatments, no

differences existed in fungal loads between the treatments (Table 5.4). This

indicates that while compost extract application may have changed fungal species

composition; there was no immediate effect on total fungal (cfu) load.

Soil temperature and soil respiration did not differ between any of the

treatments and soil types. The level of soil microbiota activity was unlikely to be

sufficiently high to raise soil temperature. Soil type (as related to solar

absorptivity) is likely to impact soil temperature, and therefore perhaps soil

respiration, but sample variability was sufficiently high to disguise any such

relationship.

150
Chapter 5: Compost extract and plant nutrition

5. Conclusions

It was hypothesised that compost extracts could increase soil

solubilisation and mineralisation rates, and thus increase plant growth. In the

experiments undertaken in this study, data suggest that the impact of compost

extract on plant nutrition and growth was apparently though the direct addition of

nutrients, rather than through an indirect effects of soil microbiota. This result is

ascribed to the high rate of extract addition, at 2 L per pot (equivalent to 34,000 L

ha-1). Compost extracts produced using the two extraction methods (with and

without aeration), did not vary in their effect on plant growth as demonstrated by

the responses of tomato and sorghum crops. Further experimentation is needed to

assess compost extract contribution to plant nutrition through increasing soil

microbiota, adding microbiota key to certain nutrient cycles.

151
Chapter 6: Compost extract and biodegradation

Chapter 6

Biodegradation of sugarcane trash through the use

of enhanced compost extracts


A revised version of this chapter has been submitted to Compost Science and Utilisation
(currently under review)

Abstract

Compost extracts were produced from rumen waste composted for one,

three, six and ten months, and from nine month old rumen compost and

sugarcane stubble (1:1 v/v mix) further composted for one month. The extracts

produced from sugarcane stubble amended compost enhanced the rate of

degradation of intact trash mat lying over a red ferrosol over a six month period

as indicated by the integral of CO2 release, with a significantly higher total

microbial activity recorded on sugarcane amended compost extract compared to

other extracts. Increased concentrations of total soil C and dissolved organic C

were found in soil samples covered with harvest residue of sugarcane compared

to the bare soil. The results of the current study indicate that compost extracts

have potential benefits to the ‗green trash‘ sugarcane production system.

152
Chapter 6: Compost extract and biodegradation

1. Introduction

Sugarcane is a crop economically important to Australia. The Australian

sugar industry is mainly concentrated along the coastal Queensland and northern

New South Wales regions. Traditional management involved burning of the

sugarcane crop before harvest, with consequent loss of dry matter and N (ranging

from 77 to 97%) and other mineral nutrients (Mitchell et al., 2000). The

alternative, ‗greentrashing‘, involves topping out the leaves from the stalk during

harvest and shredding into pieces and leaving to dry in the field (Hall et al.,

2006). The green cane trash blanketing system has been increasingly adopted by

Australian cane growers over the past two decades (Meier, 2007; Wood, 1991);

with around 70% of the sugarcane currently grown under this system (Kingston

& Norris, 2001). This approach provides mulch, resulting in reduced soil

compaction, improved soil structure, greater water and nutrient retention,

improvement of the physical and microbial status of the soil, reduction of

fertiliser demand and ultimately increased cane yield (Ball-Coelho et al., 1993;

Ball-Coelho et al., 1992; Barzengar et al., 2002; Blair & Crocker, 2000;

Casagrande et al., 1995; Graham et al., 2002; Robertson & Thorburn, 2007).

Agricultural management practices, such as stubble retention systems directly

impact microbial community structure of soil and affect the biogeochemical

processes, for example, cycling of major nutrients, rate of decomposition and

other important ecological functions (Swift & Anderson, 1994; Wakelin et al.,

2007). Trash decomposition allows for N mineralisation and incorporation of

organic C into the soil (Sutton et al., 1996; Wood, 1991). Additional benefits of

the green trash blanketing system, that attracted growers to adopt this system,

153
Chapter 6: Compost extract and biodegradation

include better weed control, reduced irrigation in dry periods, reduced soil

erosion rates and reduced labour requirement (Norrish, 1996; Small, 2000).

In the Australian sugarcane production system, green waste from the

harvest is left as mulch. It can not be incorporated into the soil if there is a rattoon

crop. The covered soils are cooler and the second (rattoon) crop is delayed. Some

disadvantages include decreased soil temperatures under the trash blanket, with

subsequent delay in ratooning, and an increase in various pests (e.g. rats) and

diseases between crops (Stewart & Wood, 1987). Rapid decomposition of the

trash is, therefore, desirable.

Soil microbiota play a central role in decomposition of organic residue in

soils, and the rate of this turnover can be increased by microbial enhancement

using microbially enhanced compost extracts (Ingham, 2005; Ryan, 2003). Some

of the fungal species known to aid in degradation of organic matter, and prevalent

in most soils, are Trichoderma spp. and cellulose digesting brown rot fungi, such

as Coniophora prasinoides, C. puteana (Highley, 1980) and Cellulomonas spp.

(Lines-Kelly, 2004). A compost extract from an inoculum source of compost

which is likely to favour proliferation of relevant organisms (cellulolytic bacteria

or fungi) may act to speed-up sugarcane trash decomposition. Some examples of

dominant bacteria, such as lactic acid bacteria, Bacillus spp., Micrococcus spp.,

Pseudomonas spp., and dominant fungi, such as Penicillium spp., Trichoderma

spp., Aspergillus spp., yeast, actinomycetes and Streptomyces spp. are reported in

microbially enhanced compost extracts (Naidu et al., 2010).

Microbial populations are likely to change during the composting process.

Presumably early successional stage composts, that are microbially active, are

needed to produce compost extracts for spraying onto the trash mat. Hall et al.

154
Chapter 6: Compost extract and biodegradation

(2006) applied extracts prepared from a compost mix of aquaculture pond

sediment, ground sugarcane, wood chips, bagasse and corn silage / cotton seed

blend over the trash mat of sugarcane. They reported increased microbial

respiration (over 168 h) of the trash mat following treatment with extracts from

eight week old compost, as compared to extracts from four, twelve and sixteen

week-old composts. This result was ascribed to an increase in bacterial

population load.

The compost mix used by Hall et al. (2006) was, however, complex. For

adoption of the use of compost extract for enhanced trash breakdown on a broad

scale basis by the sugar industry, a readily available, consistent compost

inoculum is required. Consider that if 100 L of compost extract were applied per

hectare, across the 300,000 ha of the Australia sugarcane crop, 30 ML of extract

would be required. If produced using a 1:10 water: compost mix, approximately

3,000 m3 of compost would be required annually. The beef industry overlaps

geographically with that of the sugar industry, with abattoirs located within

reasonable distance to the cane growing areas.

An experiment was, therefore, established to test the efficacy of rumen

content compost extract in degradation of harvest residue. The experiment

utilised the microbially enhanced extracts of rumen waste composted for one,

three, six, and ten months and compared these with an extract produced from a

feedstock mix of nine month old compost and the same sugarcane residue, which

was further composted for one month, and with a water control. The hypothesis

set for the possible benefits of compost extract to plants is that application of

compost extract to the sugarcane trash covered soil could enhance the

degradation of harvest residue due to the microbial action of compost extracts.

155
Chapter 6: Compost extract and biodegradation

This may subsequently increase soil C content through the addition of organic

matter via partial decomposition of plant residues.

2. Materials and methods

2.1. Soil, plant and compost material

Ferrosol soil was obtained from an organic sweet potato farm of

Rossmoya (from a depth of 30 cm) for this experiment. The principal

characteristics of this soil were previously studied and discussed in Chapter 5.

The rumen waste, obtained from an abattoir (Broadmeadows, Qld), is a

lignocellulosic by-product rich in C/N ratio.

Sugarcane trash was collected from a freshly harvested sugarcane field in

Bundaberg, Qld region and stored air dry at 4˚C. Typically, this trash contains

lignocellulosic materials and comprises of cellulose (approximately 40%),

hemicellulose (25%), lignin (18-20%) and protein (34 mg g-1 ) (Singh et al.,

2008). I chose intact over chopped trash mat for the experiment because Hall et

al. (2006) did not find any difference between the biodegradation of either

chopped or intact trash mat. Further, use of intact trash mat is more field relevant

compared to chopped or ground material.

2.2. Biodegradation of trash mat

Different aged rumen content composts from one (1MC), three (3MC), six

(6MC) and nine month old were acquired from Broadmeadows Pty Ltd,

Rockhampton. Some of the nine month old compost was further processed and

manipulated by adding sugarcane trash at the ratio of 1:1 (v/v) and allowed to

compost further for one month (10SC) in a partially enclosed condition with

156
Chapter 6: Compost extract and biodegradation

water sprinkled occasionally. The remaining portion of nine month old compost

was incubated under the same condition for a month but without the addition of

sugarcane trash (10MC) for comparison purpose.

Microbially enhanced extracts were prepared as in Chapter 2 from the five

composts. Extracts were misted over equal amounts of sugarcane trash

(approximately 100 g f wt) laid over 5 kg f wt of soil (ferrosol) in partially

enclosed plastic containers (0.28 m2 surface area and 20 L volume) in an

evaporatively-cooled greenhouse from April to October, 2009. Equal volumes of

a water-control were similarly applied. All of the treatments were applied twice at

the rate of 200 L ha-1 (i.e. 1.4 mL per container). Negative controls for all the

treatments were also established using containers with soil only. During the

experimental period, other than during respiration measurement, the containers

were covered with perforated plastic sheet to reduce water loss while maintaining

an aerobic condition. Water was sprayed over the trash or soil surface once a

month at 50 mL per pot to maintain humid conditions inside the pots. There was

no significant difference in the moisture content of soil collected from each

treatment at the end of the study (data not presented). The perforated plastic

covers were replaced with airtight plastic lids for 24 h in order to trap the CO2

produced through respiration inside the containers on days 1, 7, 14, 21, 28, 56,

84, 112, 140 and 168 of the incubation period. In terminating the experiment (168

day), the trash blanket was manually removed and weighed, allowing calculation

of weight loss. The experiment was conducted using a completely randomised

design with three replications. Total organic C, dissolved organic C and microbial

biomass C of the soil core samples collected from the top 5 cm at the end of the

157
Chapter 6: Compost extract and biodegradation

experiment were determined by the chloroform-fumigation method (Vance et al.,

1987).

2.3. Analytical methods

2.3.1 Physico-chemical analyses

The pH and EC were measured on the differently aged composts (1:20

w/v suspension), and on soil samples collected at the end of the experiment

(prepared as a 1:5 w/v suspension). Compost extracts (1:10 w/v mix) were

prepared indoors and the parameters, such as pH, EC, temperature and dissolved

oxygen were measured at 3 h intervals over a 24 h period and a final reading was

taken at 48 h, following the methods as described in detail in Chapter 2. The

moisture content of the soil samples was determined following drying of the

samples at 105˚C for 24 h. Inorganic nutrient analyses of the same composts and

compost extracts thus produced, were performed using ‗RQflex plus‘ colorimetric

test strips (Merck, Germany). All samples were analysed in triplicate.

2.3.2. Microbial analyses

Serially diluted populations of bacteria and fungi of the different aged

composts as well as compost extracts were enumerated by the pour plate count

method (Harrower, 2006). Basal and substrate induced respiration of the compost

samples were analysed following the methods as described in Chapter 3.

Microbial activities in composts, compost extracts and soil samples were assessed

by measuring fluorescein diacetate (FDA) hydrolysis following the protocol of

Adam and Duncan (2001) as described in Chapter 2.

158
Chapter 6: Compost extract and biodegradation

Microbial respiration was determined by the amount of CO2 trapped in 15

mL of 0.5 M NaOH solution inside the sealed containers over 24 h, as measured

by the acid titration method described in Chapter 3 (Anderson, 1982). Cumulative

respiration rates of sugarcane trash materials were calculated by subtracting the

respiration values generated by bare soil from those by soil with trash cover.

2.4. Statistical analyses

Data were analysed using factorial ANOVA of XLSTAT (2010), and

statistical significance of differences was evaluated at p< 0.05 using Tukey‘s test.

3. Results and discussion

3.1. Physico-chemical, biochemical and microbial analyses of different

aged composts

Samples were taken from rumen content piles that had been composting

for different periods (one, three, six and nine months), representing a pseudo-time

series of the composting process. There was a trend for pH to increase, EC to

decrease (except for an inconsistent result at nine months), NH4-N to increase to

three months, then decrease, basal respiration to decrease, and microbial activity

(FDA), bacterial plate count and fungal plate count to decrease with time of

composting (Table 6.1). The pH rise could be attributed to proteolysis as

suggested by Garcia et al. (1991). The EC decrease is attributed to leaching of the

open to atmosphere piles, and incorporation of mineral salts by microbes. NH4-N

increase to three month was consistent with proteolytic activity. The subsequent

decrease in one month old compost and nine month old compost amended with

equal volume of sugarcane trash, which was further composted for a month

159
Chapter 6: Compost extract and biodegradation

compared to other composts (Table 6.1) could be due to the presence of yet to be

mineralised organic substrates, which were highly lingo-cellulosic in nature. The

decrease in microbial activity is consistent with ‗maturation‘ of the compost

process.

Addition of sugarcane trash to nine month old rumen content compost

effectively ‗set the clock back‘ on these indices of the composting process, with

pH akin to six month, basal respiration to one month, microbial activity (FDA) to

six month, fungal plate count to one month rumen content compost, while

bacterial activity exceeded any sample of rumen content compost. The order of

magnitude of bacterial populations follows different phases of composting and

indicates that early aged compost extracts have higher metabolic activity than that

of the later phases of composting, the phases associated with the depletion of

readily available fractions of carbon from aged compost (Garcıa-Gomez et al.,

2003; Herrmann & Shann, 1997; Wu & Ma, 2002). However, addition of

sugarcane substrates in nine month old compost and further composting it for a

month must have added water soluble C and nitrogen and thus enhanced

propagation of bacteria and increased their plate counts of bacteria compared to

three, six and ten month old composts.

The basal respiration was higher in one month old compost and nine

month old compost amended with equal volume of sugarcane trash, which was

further composted for a month, while substrate induced respiration was higher in

one month old compost compared to other composts (Table 6.1). Increased

evolution of CO2 is indicative of the presence of higher populations of active

microorganisms in the immature compost (e.g. one month old compost) as

compared to the mature composts. Respiration decreases during the maturation or

160
Chapter 6: Compost extract and biodegradation

stabilisation processes (Wu et al., 2000). Increased values of bacterial and fungal

plate counts, basal respiration, and microbial activity in nine month old compost

amended with equal volume of sugarcane trash, which was further composted for

a month compared to ten month old compost (Table 6.1) might be due to the

similar reason discussed earlier.

Table 6. 1. Physico-chemical, biochemical and microbial comparisons of different aged, and


sugarcane stubble amended, composts. 1MC - one month old compost, 3MC - three month old
compost, 6MC - six month old compost, 10MC - ten month old compost, 10SC - nine month old
compost amended with equal volume of sugarcane trash and further composted for a month.
Within rows, means (n = 3) with the same letter are not significantly different according to
Tukey‘s test (p< 0.05).
Parameters 1MC 3MC 6MC 10MC 10SC
pH (H2O) 6.49 ± 0.04b 7.57 ± 0.03d 7.64 ± 0.04d 6.20 ± 0.02a 7.01 ± 0.02c
EC (dS m-1) 0.81 ± 0.01c 0.62 ± 0.01b 0.53 ± 0.01a 1.55 ± 0.01d 0.64 ± 0.00b
+
NH4 -N (mg kg-1 d wt) 63.4 ± 1.48a 676 ± 35.3d 271 ± 5.79c 273 ± 9.41c 168 ± 18.2b
Basal respiration (µg CO2 g-1 d wt h-1) 157 ± 3.80b 62.7 ± 5.17a 66.4 ± 6.27a 47.6 ± 9.10a 138 ± 10.2b
SIR respiration (µg CO2 g-1 d wt h-1) 575 ± 28.1d 290 ± 17.9b 137 ± 10.9a 456 ± 3.97c 257 ± 21.3b
Total microbial activity (µg FDA g-1 d wt h-1) 564 ± 85.5bc 853± 118.3c 518 ± 68.3b 76.3 ± 15.0a 540 ± 14.0b
Bacterial population (cfu Log10) 7.06 ± 0.04b 7.00 ± 0.03b 7.86 ± 0.01c 6.10 ± 0.10a 11.1 ± 0.01d
Fungal population (cfu Log10) 5.33 ± 0.05c 4.95 ± 0.03b 4.77 ± 0.04ab 4.55 ± 0.07a 5.52 ± 0.01c

3.2. Characterising the incubation process using different aged

compost extracts as inoculum

Dissolved oxygen declined rapidly in extracts seeded with one month old

compost and nine month old compost amended with equal volume of sugarcane

trash, which was further composted for a month, indicative of high respiratory

demand (Fig. 6.1). This finding broadly agrees with the data from the microbial

plate counts, basal respiration and total microbial activity for those composts

(Table 6.1). The pH was significantly higher in ten month old sugarcane amended

compost extract than other treatments at 24 and 48 h of incubation, however EC

of six month old compost extract and ten month old compost extract remained

higher throughout the incubation period, with few exceptions, compared to the

remaining treatments (Fig. 6.1). This agrees with the findings of Garcia et al.

161
Chapter 6: Compost extract and biodegradation

(1991) who reported higher EC in mature compost extracts compared to the

extracts of immature composts. Although temperature of ten month old sugarcane

amended compost extract increased significantly above that of other treatments at

certain times during incubation, no difference was noted between treatments at 24

and 48 h of incubation (Fig. 6.1).

162
Chapter 6: Compost extract and biodegradation

9
A
8
7
6
5

pH
4
1MCE
3 6MCE
3MCE
2 10MCE
10SCE
1
0
0 3 6 9 12 15 18 21 24 48

5.0
B
4.5
4.0
3.5
EC (dS m -1)

3.0
2.5
2.0 1MCE
6MCE
1.5 3MCE
10MCE
1.0 10SCE
0.5
0.0
0 3 6 9 12 15 18 21 24 48

35
C
30

25
Temperature (°C)

20

15
1MCE
6MCE
10 3MCE
10MCE
5 10SCE

0
0 3 6 9 12 15 18 21 24 48

D 120

100
Dissolved Oxygen (%)

80
1MCE
6MCE
60 3MCE
10MCE
10SCE
40

20

0
0 3 6 9 12 15 18 21 24 48
Time (h)

Figure 6. 1. Comparison of pH, electrical conductivity (dS m-1), fluctuations in temperatures (˚C)
and dissolved oxygen (% of saturation) over time (readings taken at every 3 h intervals until 24 h
and final readings taken at 48 h) between differently aged and sugarcane stubble amended
compost extracts. 1MCE - one month old compost extract, 3MCE - three month old compost
extract, 6MCE - six month old compost extract, 10MCE - ten month old compost extract, 10SCE
- compost extract produced from nine month old compost amended with equal volume of
sugarcane trash and further composted for a month. Values are mean ± SE (n = 3).

163
Chapter 6: Compost extract and biodegradation

3.3. Comparison of chemical, microbial and biochemical

characteristics of different aged and sugarcane stubble amended

compost (24 h incubated) extracts

Plate counts of compost extracts revealed that bacterial populations were

found in the order; one month old compost extract > ten month old sugarcane

amended compost extract > three month old compost extract > six month old

compost extract > ten month old compost extract (Table 6.2), consistent with the

bacterial load of the inoculum composts. This indicates that the incubation

conditions supported bacterial growth. In contrast, there was little difference

between the treatments in terms of culturable fungal load, despite differences in

the composts per se, suggesting that the incubation conditions did not favour

fungal growth.

No differences were found in phosphate, potassium and fulvic acids, but

the ammonium and humic acid concentrations were found to be higher in ten

month old sugarcane amended compost extract compared to all other compost

extracts (Table 6.2). These increments must have been sourced from sugarcane

stubble added to the nine month old compost. Garcia et al. (1991) found

significant reduction in NH4+-N during composting and maturation, however

ammonium concentration increased initially then declined after three months of

composting (Table 6.2), the reason for which is unclear.

164
Chapter 6: Compost extract and biodegradation

Table 6. 2. Comparison of chemical, biochemical and microbial characteristics of different aged


and sugarcane stubble amended compost extracts incubated for 24 h with 1% kelp and 0.5%
molasses as amendments. 1MCE - one month old compost extract, 3MCE - three month old
compost extract, 6MCE - six month old compost extract, 10MCE - ten month old compost extract,
10SCE - compost extract produced from nine month old compost amended with equal volume of
sugarcane trash and further composted for a month. Within rows, means (n = 3) with the same
letter are not significantly different according to Tukey‘s test (p< 0.05).
Parameters 1MCE 3MCE 6MCE 10MCE 10SCE
+
NH4 -N (mg L-1) 0.54 ± 0.04a 0.91 ± 0.05b 0.67 ± 0.05a 0.65 ± 0.03a 1.11 ± 0.03c
PO43--P (mg L-1) 25.7 ± 1.09a 22.0 ± 1.43a 22.0 ± 1.70a 22.0 ± 0.58a 25.7 ± 1.15a
K+-K (g L-1) 0.57 ± 0.06a 0.58 ± 0.01a 0.63 ± 0.02a 0.62 ± 0.03a 0.56 ± 0.02a
Humic acid (g L-1) 0.17 ± 0.01a 0.22 ± 0.05a 0.13 ± 0.03a 0.18 ± 0.03a 0.41 ± 0.01b
Fulvic acid (g L-1) 0.23 ± 0.01a 0.26 ± 0.03a 0.28 ± 0.00a 0.28 ± 0.00a 0.24 ± 0.00a
Bacterial population (cfu log 10) @ 24 h 12.2 ± 0.02e 11.1 ± 0.02c 10.5 ± 0.03b 10.3 ± 0.02a 11.4 ± 0.01d
Fungal population (cfu log 10) @ 24 h 5.63 ± 0.03b 5.70 ± 0.09b 5.20 ± 0.04a 5.54 ± 0.06b 5.45 ± 0.04ab

The least microbial activity in ten month old compost extract as compared

to one, three and six month old compost extracts suggests that active microbiota

decline as the compost matures (Fig. 6.2). Nevertheless, the addition of sugarcane

trash in nine month old compost greatly stimulated and revived the microbial

activity of the resultant compost extract (Fig. 6.2). Higher microbial activity in

compost extract produced from ten month old sugarcane amended compost

extract as compared to ten month old compost extract implies that the addition of

sugarcane substrate might have supplied energy (in the forms of the labile and

total organic C) to the less active microbiota of the ten month old compost. This

could be related to a priming effect or substrate stimulation causing

mineralisation of organic matter, as found with the addition of glucose to soil

(Shen & Bartha, 1996).

165
Chapter 6: Compost extract and biodegradation

Figure 6. 2. Comparison of total microbial activity at different times (readings taken at 12, 24 and
48 h) following the FDA hydrolysis method in different aged (one, three, six and ten month old)
and sugarcane stubble amended compost extracts. 1MCE - one month old compost extract, 3MCE
- three month old compost extract, 6MCE - six month old compost extract, 10MCE - ten month
old compost extract, 10SCE - compost extract produced from nine month old compost amended
with equal volume of sugarcane trash and further composted for a month. Within a given time,
treatment means with the same letter do not differ significantly and values are mean ± SE (n = 3).

When solid composts were compared with liquid forms of compost

extracts incubated with microbial food sources and continuous supply of aeration,

the bacterial populations increased tremendously in the extracts, irrespective of

the treatment effects (compare data in Tables 6.1 and 6.2). Given the bacterial

counts of ten month old compost (1.3 x 106 cfu g-1) and ten month old compost

extract (2.1 x 1010 cfu mL-1), there was a significant increase in the number of

bacteria when transformed from solid to liquid form. Similarly, fungal

populations were slightly increased by liquid incubation in all treatments with an

exception of ten month old sugarcane amended compost extract, but not by as

much as bacteria. Total microbial activity (FDA hydrolysis) likewise increased

by varying amounts in liquid extracts compared to their original forms, except in

six month old compost extract which was similar in both solid and liquid forms

(Table 6.1 and Fig. 6.2). These results suggest that conversion of solid composts

166
Chapter 6: Compost extract and biodegradation

into liquid forms under the stated conditions is beneficial for microbial

enrichment in the resulting extracts.

It seems that the benefit of compost extract over compost is in the

multiplication of microbes. The next question is survival of these microbes in the

soil. For example, rhizobia are liquid-cultured and can be applied directly to plant

roots/soil. But in the field, survivability of these microbes is low (storage plus in-

field survival). Therefore, a delivery system, such as peat is normally used.

Nevertheless, previous studies have shown that rhizobia was successfully

delivered into the soil via drip irrigation system (biofertigation) compared to a

conventional seed pelleting system (Mussaddak & Fawaz, 2008). Inoculation of a

liquid culture of rhizobia through this system showed enhanced N2 fixation by

soybean plants. This system could also apply to compost extracts as they deliver

beneficial microorganisms into a field soil.

3.4. Effects of compost extract treatments on sugarcane trash

degradation

After 168 d, soil pH was lower in treatments involving cane trash (Table

6.3) which is in accordance with the findings of Robertson (2003). There was no

difference in soil pH between treatments receiving different compost extract but

without trash, in contrast to when trash was present. Although no difference was

noted between the non-covered bare soil treated with different compost extracts,

total and dissolved organic C were higher in soils covered with sugarcane trash

(Table 6.3). Mulching generally increases microbial load and respiration in soil as

indicated by the release of CO2-C (measured with fumigation extraction) and by

the stimulation of substrate induced respiration (SIR), respectively (Wardle et al.,

167
Chapter 6: Compost extract and biodegradation

1999). For microbial biomass C, there was only one major difference between

treatments. There was a significant increase in microbial biomass C in ten month

old sugarcane amended compost extract applied to bare soil as compared to the

water-control treatment with sugarcane stubble (Table 6.3). The literature shows

that numerous factors, such as soil-trash contact (Magid et al., 2006), and climate

and soil interaction (Young & Ritz, 2000) influence increases of microbial

biomass (Robertson & Thorburn, 2007).

There was no significant difference in total microbial activities of soil

samples collected across treatments either with or without a trash cover, however

soil microbial activity was comparatively higher (two-fold or more), though not

at significant level, when sugarcane stubble was retained compared to the bare

soil (Table 6.3). Higher microbial activity (as shown by the respiration data) in

the compost extract treated trash indicated that once the microbes are able to

degrade the lingo-cellulosic secondary cell walls, the decomposition should

proceed quickly. A reason for lower soil biological activity without trash is

generally linked to the decline in dissolved organic C (labile C) (Bell et al.,

2007), which is in line with the findings of the current study.

168
Chapter 6: Compost extract and biodegradation

Table 6. 3. Chemical, biochemical and microbial characteristics of soil samples (collected after 168 days of incubation) covered with or without sugarcane trash and treated
with compost extracts produced from differently aged and sugarcane stubble amended compost and incubated for 24 h with 1% kelp and 0.5% molasses as amendments.
1MCE - one month old compost extract, 3MCE - three month old compost extract, 6MCE - six month old compost extract, 10MCE - ten month old compost extract, 10SCE
- compost extract produced from nine month old compost amended with equal volume of sugarcane trash and further composted for a month. Within rows, mean (n = 3) ±
SE with the same letter are not significantly different according to Tukey‘s test (p< 0.05).

with trash without trash


Parameters Cont 1MCE 3MCE 6MCE 10MCE 10SCE C Cont C 1MCE C 3MCE C 6MCE C 10MCE C 10SCE
pH (H2O) 5.10 ± 0.10a 5.51 ± 0.01cd 5.53 ± 0.02cd 5.44 ± 0.02bc 5.46 ± 0.04bc 5.21 ± 0.12ab 5.56 ± 0.04cd 5.68 ± 0.02cd 5.70 ± 0.02cd 5.75 ± 0.01d 5.64 ± 0.05cd 5.70 ± 0.01cd
EC (dS m-1) 0.11 ± 0.00ab 0.14 ± 0.00b 0.12 ± 0.00ab 0.11 ± 0.01a 0.12 ± 0.01ab 0.11 ± 0.01ab 0.12 ± 0.00ab 0.11 ± 0.01ab 0.11 ± 0.00ab 0.11 ± 0.00a 0.11 ± 0.01ab 0.11 ± 0.01ab
Total microbial activity (µg FDA g-1 d wt h-1) 674 ± 70.8a 732 ± 35.5a 589.2 ± 34.9a 489 ± 331.4a 553 ± 228a 723 ± 73.6a 276 ± 28.7a 207 ± 12.8a 267 ± 40.7a 224 ± 12.0a 266 ± 58.1a 401 ± 82.7a
-1
Total organic C (mg C g d wt) 0.07 ± 0.00c 0.07 ± 0.00c 0.08 ± 0.00c 0.08 ± 0.00c 0.08 ± 0.00c 0.08 ± 0.00c 0.02 ± 0.00a 0.04 ± 0.01b 0.03 ± 0.00ab 0.02 ± 0.00a 0.02 ± 0.00a 0.03 ± 0.00a
-1
Dissolved organic C (mg C g d wt) 0.07 ± 0.00b 0.06 ± 0.00b 0.06 ± 0.01b 0.07 ± 0.00b 0.07 ± 0.00b 0.07 ± 0.01b 0.01 ± 0.00a 0.03 ± 0.00a 0.01 ± 0.00a 0.01 ± 0.00a 0.01 ± 0.00a 0.01 ± 0.00a
-1
Microbial biomass C (mg C g d wt) 0.02 ± 0.01a 0.03 ± 0.00ab 0.05 ± 0.01ab 0.04 ± 0.00ab 0.02 ± 0.00a 0.05 ± 0.01ab 0.03 ± 0.00ab 0.05 ± 0.03ab 0.04 ± 0.00ab 0.03 ± 0.01ab 0.04 ± 0.00ab 0.09 ± 0.03b
Trash weight (g f wt) 98.2 ± 0.88a 95.2 ± 5.64a 98.0 ± 2.02a 97.5 ± 2.29a 95.2 ± 2.46a 90.8 ± 1.17a N/A N/A N/A N/A N/A N/A
N/A - Not applicable

169
Chapter 6: Compost extract and biodegradation

3.5 Weight loss of trash and cumulative respiration rate

Although ten month old sugarcane amended compost extract showed the

maximum loss in weight (f wt) of sugarcane trash, there was no significant

difference among the treatments (Table 6.3). In contrast, Witkamp (1966) found a

significant loss in weight (i.e. 40-90%), using a mesh-bag technique, of four

tested leaf litters (mulberry, redbud, oak and pine) within a year under a natural

forest environment. This is logical considering the one-year long trial period

compared to the current six month biodegradation trial. This weight loss of

various leaf species in their study was attributed to the leaching of water soluble

nutrients from their surfaces and was highly correlated with the microbial load

present in the litter. In contrary, perhaps there was no loss of soluble nutrients

from the surfaces of sugarcane litter in the present study.

In the current experiment, however, sugarcane trash treated with compost

extracts did show visual signs of degradation, such as discolouration, compared

to those treated with water only (Fig. 6.3). This could be due to leaching of

nutrients caused by the mineralisation of organic matter of trash or due to

degradation caused by microbial attack or cellulose utilisation by the microbes

(Parsons & Congdon, 2008).

170
Chapter 6: Compost extract and biodegradation

Figure 6. 3. On the left (A) showing sugarcane trash treated with water control and on the right
(B) showing the same trash treated with compost extract produced from nine month old compost
amended with equal volume of sugarcane trash and further composted for a month (10SCE).

Although there was no significant weight loss by ten month old sugarcane

amended compost extract, cumulative respiration as measured in the ten month

old sugarcane amended compost extract treated sugarcane trash was significantly

higher (p< 0.05) from day 1 to day 56 compared to that of the trash control

treatment, after which no significant difference was noted in the cumulative

respiration between treatments (Fig. 6.4). This could be due to the loss of

moisture inside the containers in the due course of time that there was no

difference in weight loss of sugarcane trash between treatments. It could also be

because the time period was insufficient to allow a difference in trash weight to

be expressed. Indeed the cumulative respiration of 230000 ug CO2 pot-1 is

equivalent to 230 x 12 / 44 = approximately 63 mg C pot-1 lost in respiration, a

trivial amount compared to the trash weight of approximately 95 g pot-1.

Higher cumulative respiration from ten month old sugarcane amended

compost extract treated sugarcane trash could possibly be due to addition and

stimulation of microbiota present in the sugarcane layer or the soil through the

application of compost extracts. Another possible reason for this could simply be

due to the combined synergistic effect of soil indigenous and exogenous

171
Chapter 6: Compost extract and biodegradation

microbiota, the latter applied in the form of organic amendments or biofertilisers

(Prasanna et al., 2008). Synergistic or antagonistic effects of microbial diversity

in litter decomposition of the terrestrial ecosystems have been reviewed by

Hattenschwiler et al. (2005). Synergistic interaction, in general, enhances the rate

of decomposition unlike the antagonistic interaction, in which microorganisms

compete for the similar resources, thereby slowing the rate of litter decay.

Nevertheless, no such antagonistic effect has been noticed in the current study

because there was a higher cumulative respiration in each treatment compared to

that of the water control.

172
Chapter 6: Compost extract and biodegradation

Figure 6. 4. (A) Time course of respiration rates which was assessed as accumulation of CO2
over a 24 h period and (B) comparison of cumulative respiration rates of sugarcane trash materials
lying on top of the ferrosol soil over six month (28 weeks) from 22nd April to 28th October, 2009
(readings taken at 1, 7, 14, 28, 56, 84, 112, 140 and 168 days of incubation period) applied with
different aged and sugarcane stubble amended compost extracts as measured by trapping CO2
evolved. Respiration rates and cumulative respiration rates of sugarcane trash materials were
calculated by subtracting the respiration values generated by bare soil from those by soil with
trash cover. 1MCE - one month old compost extract, 3MCE - three month old compost extract,
6MCE - six month old compost extract, 10MCE - ten month old compost extract, 10SCE -
compost extract produced from nine month old compost amended with equal volume of sugarcane
trash and further composted for a month. Values are mean ± SE (n = 3).

There was distinct seasonal variation in microbial respiration; during the

winter (weeks 8 to 16), the rate of respiration was much reduced compared to the

warmer summer (Fig. 6.4) although the containers were kept in the greenhouse

(i.e. in a semi-controlled environment). Other studies have also shown a direct

influence of environmental variables, for example, ambient temperature and

moisture, on soil respiration (Bingrui & Guangsheng, 2009; Han et al., 2007;

173
Chapter 6: Compost extract and biodegradation

Lloyd & Taylor, 1994) and ultimately on the rate of decomposition and nutrient

cycling (Davidson et al., 2000; Heal et al., 1997; Zak et al., 1999).

4. Conclusions

Microbially enhanced compost extract produced from a mixture of

composted rumen waste and sugarcane stubble has the potential to accelerate

degradation of a sugarcane trash mat, as indicated by the enhanced cumulative

respiration in this study. It is evident that none of the differently aged compost

extracts enhanced degradation of an intact trash mat, unless sugarcane stubble

was incorporated into the compost. This suggests that incorporation of the harvest

residue into the rumen content compost selected for microorganisms specific for

degradation of cane trash. Increased levels of organic C in soil covered with

harvest residue, irrespective of the compost extract treatments, highlighted the

benefits of mulching. Future research should focus on optimising conditions for

degradation, such as comparing trash particle size and incorporation of trash into

soil, besides the use or not of compost extract in situ.

174
Chapter 7: Compost extract and disease suppression

Chapter 7

Suppressing fungal wilt diseases of tomato caused

by Fusarium oxysporum, F. solani and Rhizoctonia

solani using compost extracts in vitro and in

hydroponics culture
A revised version of this chapter has been accepted in Biological Control, prepared by Karuna
Shrestha, Pramod Shrestha, Kerry B. Walsh, Keith M. Harrower and David J. Midmore

Abstract

The efficiency of compost extracts prepared aerobically or anaerobically,

with three or nine month old compost, and subsequently sterilised or not, on

disease suppression of pathogenic fungi was studied. Fresh, but not sterilised,

compost extracts suppressed mycelial growth of three fungal pathogens, namely,

Fusarium oxysporum, Fusarium solani, and Rhizoctonia solani by 82, 79 and

88%, respectively, compared to the non-treated control. The highest inhibition

was demonstrated by aerated extract based on three month composted material.

Sporulation of F. oxysporum and F. solani was also reduced by all non-sterilised

compost extracts aged three and nine months, with overall spore reduction by up

to 98%. Sterilised compost extracts enhanced mycelial growth and sporulation in

most cases. Hydroponically cultured tomato plants (cv. Tiny Tim) inoculated

with conidial suspensions (1 x 106 spores per plant) of the pathogenic fungi F.

175
Chapter 7: Compost extract and disease suppression

oxysporum had less severe fungal wilt symptoms when compost extract was

added to the medium, however disease suppression on inoculation with F. solani

and R. solani was not significantly reduced. In the hydroponics studies, there was

little difference in terms of benefit between aeration of compost extracts, or age

of composts.

1. Introduction

Fungal pathogens, such as Fusarium oxysporum, Fusarium solani, and

Rhizoctonia solani are important plant pathogens, which cause severe damage to

several fruit and vegetable crops, including tomato. Control of these

phytopathogenic fungi typically relies on use of the synthetic fungicides Thairam,

Dichlofluanid, and Fluazinam (Elad et al., 2007). These fungicides are persistent,

and have a level of toxicity to wild life and humans (Crnko et al., 1992). Their

widespread use is implicated in the development of pathogen resistance to these

pesticides (Elad et al., 1992), thereby threatening the stability of crop production.

Elevated concentrations can cause reduced soil biological activity in agricultural

lands (Dumestre et al., 1999), and in addition, these biocides are not confined to

target species; their application can also negatively impact beneficial organisms

(Zwieten et al., 2007). Therefore, alternatives to these are extensively being

explored, so avoiding use of synthetic fungicides in order to sustainably produce

safe food.

Organic soil amendments, such as compost have been reported to be

effective in controlling several fungal diseases, such as Sclerotium rolfsii (Danon

et al., 2007; Hadar & Gorodecki, 1991), Fusarium spp. (Chef et al., 1983;

Cotxarrera et al., 2002; Henis et al., 1984) and Rhizoctonia solani (Nelson &

176
Chapter 7: Compost extract and disease suppression

Hoitink, 1983). Suppressive activities against various pathogens are considered to

be due to antagonism by beneficial microflora (Hoitink et al., 1997), including

strains of Trichoderma spp. (Cotxarrera et al., 2002), Bacillus subtilis (Phae et al.,

1990), or due to the production and release of allelochemicals (Bailey &

Lazarovits, 2003).

Compost extract is likewise reported to suppress the root diseases in

economically important crops, including tomato, caused by fungal pathogens, for

example, F. oxysporum f. sp. lycopersici (Hibar et al., 2006; Ma et al., 2001b;

Utkhede & Koch, 2004). Studies evaluating the effects of various compost

extracts on Pythium debaryanum, Fusarium oxysporum f. sp. lycopersici and

Sclerotium bataticola carried out in situ and in vitro showed that microflora

present in compost extract suppressed the growth of tested fungi by parasitism

(El-Masry et al., 2002). Research studies conducted in vitro demonstrated up to

98% inhibition of Venturia inequalis conidia germination by the antibiotic effect

of compost extract as compared to the water control (Cronin et al., 1996). The

antagonistic action of the microorganisms present in compost extracts against

pathogenic fungi may be due to direct competition for nutrients and space (Al-

Mughrabi et al., 2008), production of antimicrobial compounds or induced

systemic resistance in plants (Zhang et al., 1998). Some researchers have

suggested that the antagonistic effect may be due to compounds in the compost

extract (Hoitink et al., 1997; Siddiqui et al., 2008b).

Other studies have also reported an effectiveness of compost extracts in

suppressing Fusarium wilt of sweet pepper (Fusarium oxysporum f. sp.

vasinfectum) by up to 89% (Ma et al., 2001b). Similarly, an approximately 85%

reduction in Choanephora wet rot severity in a field situation was demonstrated

177
Chapter 7: Compost extract and disease suppression

on okra when treated with Trichoderma-fortified extract produced from a mixture

of rice straw and composted oil palm inflorescence stalks. The extract showed a

comparable effect to the conventional fungicide Dithane M-45R (Siddiqui et al.,

2008b). Scheuerell and Mahaffee (2004) reported that aerated compost extract

amended with kelp and humic acid was more effective than non-aerated compost

extract in controlling damping-off in cucumber seedlings caused by Pythium

ultimum. In contrast, some studies has shown that compost extracts produced

with continuous aeration fail to suppress fungal diseases, such as grey mold

(Botrytis cinerea) on geranium (Elad & Steinberg, 1994) and late blight

(Phytopthora infestans) on potato (Olanya & Larkin, 2006). Cronin et al. (1996)

also showed that aeration decreased the efficacy of compost extracts compared to

the non-aerated controls and ascribed this to the possible production of major

inhibitory metabolites in non-aerated extracts, which are heat stable, non-protein

in nature, and low in molecular weight. Nevertheless, Scheuerell (2003) did not

find any difference between aerated and non-aerated compost extracts in

controlling fungal diseases, both were effective against grey mold on greenhouse

grown geraniums, damping-off of basil seedlings.

In vitro studies have also been employed to test the effect of compost

extracts on fungal growth. El-Masry et al. (2002) reported that compost extracts

based on fruit waste, garden waste and crop waste suppressed growth of various

fungal species, including Fusarium oxysporum f. sp. lycopersici. Compost

extracts at concentrations of 5%, 10% and 15% (v/v) to PDA-Sigma growth

medium suppressed the hyphal growth of F. oxysporum f. sp. lycopersici by

94.4%. However, there was no antagonistic effect against the tested fungi using

autoclaved compost water extracts.

178
Chapter 7: Compost extract and disease suppression

McQuilken et al. (1994) also noted that filter sterilisation or autoclaving

of compost extracts negated the disease suppression effect, in their case against

the plant pathogen Botrytis cinerea. Furthermore, stimulatory effects of sterilised

compost extracts have been reported by Hardy and Sivasithamparam (1991) on

sporangia formation of Phytopthora spp. while non-sterilised extracts showed

inhibitory effects and induced lysis of sporangia.

Based on previous work I, therefore, expect compost extract to suppress

fungal pathogen growth, but not all extracts may have the same effect due to

different forms of compost extracts, their being microbially enhanced, aerated,

non-aerated, different aged and so on.

The aim of this study was to compare the disease suppressive effects of

compost extracts from different aged composts in vitro and in situ, and to assess

their ability to reduce disease severity on a tomato crop caused by Fusarium

oxysporum f. sp. lycopersici, F. solani DAR 66102 and Rhizoctonia solani DAR

61830. The hypothesis to be tested is that compost extract suppresses fungal root

diseases of tomato, reducing mycelial growth, spore production and total biomass

of these pathogens through its biological and/or non-biological action, thus

offering crop protection against these fungal diseases.

2. Materials and methods

The efficacy of compost extracts in suppressing fungal wilt diseases of

tomato was tested in vitro, in terms of fungal hyphal growth (trial 1and 2), and in

hydroponics culture, in terms of tomato root disease expression (trial 3 and 4).

For trials 1 and 3, Fusarium oxysporum f. sp. lycopersici (sourced from

CQUniversity, Qld) was employed.

179
Chapter 7: Compost extract and disease suppression

For trials 2 and 4, the same Fusarium oxysporum f. sp. lycopersici, and

two additional fungal pathogens: Fusarium solani DAR 66102 and Rhizoctonia

solani DAR 61830 (both sourced from Department of Primary Industries, NSW),

were used, as causal agents of fungal wilt of tomato. These strains were cultured

on Petri plates using a basal media (Harrower, 2006), consisting of Na2HPO4

0.725, KH2PO4 0.725, MgSO4.7H2O 0.12, NaCl 0.1, NH4NO3 0.4, glucose 1.8,

yeast extract 1.1, malt extract 1.0 and agar 15.0 in g per litre of water. In both

experiments, the plates were incubated at 25˚C until heavy sporulation was

observed and then stored at 4˚C until use.

Compost extract was produced by incubating nine month old rumen

compost in water at the ratio of 1: 10 (w/v), both with continuous aeration for 24

h and without aeration. The process followed was as in Chapter 2. All compost

extracts in both experiments were amended with 1% ‗fish and kelp hydrolysate‘

from Searle Pty Ltd, Australia, and 0.5% molasses at the onset of incubation

(Shrestha et al., 2011b).

2.1. In vitro hyphal growth trials

Two in vitro trials were undertaken. In the first trial, two concentrations

(50 ml L-1 and 100 ml L-1) of aerated and non-aerated extracts prepared from nine

month old rumen content compost were compared with the same concentrations

of sterilised aerated and sterilised non-aerated compost extracts and the control

(equal volumes of distilled water). In trial 2, three and nine month old compost

extracts were produced with continuous aeration and nine month old compost

extract was also produced without aeration, as described above. Sterilisation of

one half of each compost extract was achieved by autoclaving the extracts at

180
Chapter 7: Compost extract and disease suppression

121˚C for 15 minutes at 15 psi (0.1034 MPa) pressure followed by filtration

through 0.45 µm-pore membrane filters (Millipore, Australia) to assure the

absolute elimination of live microorganisms. Three parameters, namely, mycelial

growth, spore count and total biomass were assessed to evaluate the efficacy of

compost extracts in suppressing the fungi in both in vitro trials.

In both trials, aliquots of 100 mL of non-sterilised and sterilised compost

extracts and a distilled water control were added and blended thoroughly with the

melted basal medium (1 L) before pouring into plates. Following solidification, a

one cubic mm agar block with cultured fungal pathogen was placed at the centre

of each plate. Plates were then incubated at ± 25˚C for a week. Five (trial 1) and

four (trial 2) replicates were established for each treatment. Plates were randomly

placed within an incubator. Mycelial growth rate was quantified by measuring the

radial extension of the mycelia in millimeters at six days (in trial 1), and three and

six days (in trial 2) after plate inoculation, before growth reached the edge of the

Petri plates (Harrower, 2006).

Following a week after the measurement of radial extension at six days

after plate inoculation in both experiments, spores produced on each Petri plate

were collected by swirling 6 mL of distilled water with 0.1% Tween 80 and

0.01% chloramphenicol. The collected spore suspensions were stained separately

with two drops of lactophenol cotton blue dye per container. Due to the presence

of phenol in the stain, the microorganisms were killed. Spore density was

assessed by microscopic observation using a haemocytometer.

To determine the fungal biomass, in trial 1, 200 μL of spore suspension of

F. oxysporum was inoculated into 70 mL liquid basal medium contained in Schott

bottles. In trial 2, the same volume of spore suspensions of three fungal

181
Chapter 7: Compost extract and disease suppression

pathogens was inoculated into 35 mL liquid basal medium contained in Falcon

tubes under the same conditions as described for trial 1. At the same time of

pathogen inoculation, compost extracts at the rate of 50 mL L-1 and 100 mL L-1

(in trial 1) and 100 mL L-1 (in trial 2) were added to the liquid basal media, and

the mixtures were shaken constantly at 150 rpm for two weeks in the laboratory

at room temperature. The cultures were then harvested, oven dried at 55˚C for 72

hours and weighed to obtain mean mycelial biomass (d wt) in either trial. Each

treatment was replicated five and four times for trials 1 and 2, respectively.

2.2. Hydroponics trials

A non-recirculating hydroponics system (Midmore & Wu, 1999) with

eight plants per styrofoam hydroponics container was established. Tomato seeds

(cultivar Tiny Tim) were sown in vermiculite and two weeks later tomato

seedlings were transplanted individually into perforated poly-pots (3.5 cm

diameter) containing pieces of nylon net at the bottom and filled with vermiculite

to support the plant.

The pots containing the tomato seedlings were inserted into holes in the

styrofoam lids of the hydroponics containers, which were filled with reverse

osmosis (RO) water with (in trial 3) and without (in trial 4) nutrients.

Hydroponics nutrients (Manutec Pty Ltd, Australia) was used in trial 3 at the rate

of 120 g (part I) plus 80 g (part II) per 100 litre of water. Nutrient addition was

avoided in trial 4 because it could mask the effect of compost extracts.

A month after transplanting, tomato roots were inoculated with spore

suspensions using a plastic sprayer with a nozzle at the rate of 1 x 106 spores per

plant (Ramsay et al., 1992) of F. oxysporum in trial 3 and of the three pathogenic

182
Chapter 7: Compost extract and disease suppression

strains in trial 4. A second inoculation was followed a week later. As these fungal

wilt diseases are prevalent in warm weather, air temperature inside the

greenhouse was increased to 28˚C in shaded conditions in both experiments. The

pH of the solution culture was also adjusted to 4.5 by adding either HCl or

NaOH.

The compost extract treatment involved approximately 10 mL per

container, equivalent to an application rate of 725 L ha-1. The treatments were

first applied four weeks after transplanting, just prior to the first inoculation of the

tested pathogens and then secondly a week later.

A completely randomised design was used in both experiments. In trial 3,

there were three treatments, namely, control (water only), aerated compost

extract, and sterilised aerated compost extract, all with and without pathogen

inoculation to hydroponics tomato plants. Each treatment was applied to three

replicate containers, each container holding eight plants. In trial 4, six treatments

were employed: nine month old aerated compost extract (9M ACE) and nine

month old sterilised aerated compost extract (9M SACE), nine month old non-

aerated compost extract (9M NCE) and nine month old sterilised non-aerated

compost extract (9M SNCE), and three month old aerated compost extract (3M

ACE) and the water control. Each treatment was applied to three replicate

containers, each container holding eight plants. As no difference was noted in

growth of tomato plants between pathogen inoculated and non-inoculated plants

in trial 3, only pathogen inoculated plants were used in trial 4.

The presence of F. oxysporum in root tissue was assessed in trial 3. Root

samples were collected (n = 3 per treatment), and surface sterilised in a 10%

bleach solution for 30 seconds. This was followed by incubation of root samples

183
Chapter 7: Compost extract and disease suppression

(1 cm long segments) for 48 hours at 25˚C on Petri plates containing a selective

culture media. The selective culture media ‗malachite green agar‘, suitable for

isolating F. oxysporum (Castella et al., 1997), was prepared by mixing peptone (1

g), KH2PO4 (1 g), MgSO4.7H2O (0.5 g), malachite green oxalate (2.5 mg), agar

bacteriological (20 g) and malachite green agar (2.5 ppm) in one litre of RO

water. This stimulated production of fungal spores from conidia and fungal

invasion from inside of the roots, and the presence or absence of the pathogen in

the tomato root samples was noted.

In both trials, disease severity was assessed a week after the second

inoculation of pathogen in the plants, with a disease severity rating of 0-5

(Ramsay et al., 1992) (0 - no wilt, healthy plant; 1 - slightly stunted, <5% leaves

affected; 2 - slight yellowing and wilting, <25% leaves affected; 3 - moderate

wilting, 25-50% leaves affected; 4 - severe wilting, 50-75% leaves affected; and

5- >75% leaves wilted). Leaf chlorosis (yellowing) is a major symptom of wilt

disease; therefore, chlorophyll concentration of the leaves was also measured in

trial 4 as an index of disease severity using a SPAD-502 meter (Minolta

Corporation, Japan), a meter measuring absorbance at 663 nm (chlorophyll a) and

645 nm (chlorophyll b).

2.3. Statistical analyses

In both experiments, data were analysed using the GLM procedure of

XLSTAT (2010) and statistical significance was evaluated at p< 0.05.

184
Chapter 7: Compost extract and disease suppression

3. Results and discussion

3.1. In vitro trials

3.1.1. Mycelial growth of fungi

In trial 1, the effect of two concentrations (50 mL L-1 and 100 mL L-1) of

sterilised and non-sterilised aerated and non-aerated compost extracts of nine

month old compost were compared to a water control, in terms of impact on

mycelial growth of F. oxysporum f. sp. lycopersici. The mycelial growth rate of

F. oxysporum (p< 0.001) was markedly suppressed (32.8%) by the non-sterilised

aerated compost extract treatment at 100, but not 50 mL L-1 nutrient agar after six

days of inoculation as compared to the control (Fig. 7.1). Therefore, the

application rate of 100 mL L-1 was chosen for trial 2.

El-Masry et al. (2002) reported that concentration of 5% (v/v), which is

equivalent to 50 mL L-1, compost extract based on crop compost to PDA-Sigma

medium suppressed the mycelial growth of Sclerotium bataticola by 83%.

Similarly, concentration of 10% (v/v) and 15% (v/v) compost extracts based on

leafy fruit compost or garden compost suppressed the mycelial growth of the

same pathogen by 94%. In the same experiment, they found that the mycelial

growth of F. oxysporum was suppressed by 94% using either concentration or

compost source.

185
Chapter 7: Compost extract and disease suppression

Figure 7. 1. Effect of compost extracts in trial 1 on mycelial growth rates (mm d -1) of Fusarium
oxysporum, f. sp. lycopersici six days after inoculation. Cont - control (water), S ACE1 - sterilised
aerated compost extract 50 mL L-1, S ACE2 - sterilised aerated compost extract 100 mL L-1, S
NCE1 - sterilised non-aerated compost extract 50 mL L-1, S NCE2 - sterilised non-aerated
compost extract 100 mL L-1, ACE1 - aerated compost extract 50 mL L-1, ACE2 - aerated compost
extract 100 mL L-1, NCE1 - non-aerated compost extract 50 mL L-1, NCE2 - non-aerated compost
extract 100 mL L-1. Means with different letters indicate significant difference (p< 0.05) across
treatments and vertical bars represent standard errors of the means (n = 5).

In trial 2, all ‗live‘ compost extracts significantly inhibited the radial

growth of the three tested pathogens, compared to the water control. The mycelial

growth (six days after inoculation) was inhibited by approximately 82%, 79% and

88%, respectively for F. oxysporum, F. solani and R. solani on addition of three

month old extract (Fig. 7.2). In accord with these results, Hibar et al. (2006)

reported an inhibition (up to 70%) of mycelial growth of F. oxysporum f. sp.

radicis-lycopersici incubated at 25˚C for six days by addition of compost extract

produced at 1:1 (v/v) compost to water ratio and applied at the rate of 10 mL per

100 mL (which is equivalent to 100 mL L-1) of potato dextrose broth.

No significant inhibition was evident on the mycelial radial extension of

F. oxysporum or R. solani with autoclaved and microfiltered compost extracts,

indicating that the disease suppression effect is linked to the living microbiota.

This is consistent with the findings of McQuilken et al. (1994) and El-Masry et

186
Chapter 7: Compost extract and disease suppression

al. (2002) who showed a loss of antagonistic effect in autoclaved compost

extracts against the tested phytopathogenic fungi (Botrytis cinerea, Fusarium

oxysporum, Sclerotium bataticola, Pythium debaryanum). Indeed, mycelial

growth of F. solani was enhanced dramatically with both sterilised extracts (Fig.

7.2).

Figure 7. 2. Effect of compost extracts in trial 2 on mycelial growth rates (mm d -1) of Fusarium
oxysporum, f. sp. lycopersici, F. solani DAR 66102, and Rhizoctonia solani DAR 61830 over (A)
three and (B) six days after inoculation. Cont - control (water), 3M ACE - three month old
aerated compost extract, 9M ACE - nine month old aerated compost extract, 9M NCE - nine
month old non-aerated compost extract, 9M SACE - nine month old sterilised aerated compost
extract, 9M SNCE - nine month old sterilised non-aerated compost extract. Means with different
letters indicate significant difference (p< 0.05) across a given species and vertical bars represent
standard errors of the means (n = 4).

187
Chapter 7: Compost extract and disease suppression

3.1.2. Spore count

Consistent with the mycelial growth results, the sporulation of F.

oxysporum was decreased (p< 0.05) when the agar medium was amended with

either aerated or non-aerated compost extracts applied at the rate of 100 mL L-1,

in comparison to the 50 mL L-1 treatment which had much less suppressive effect

(Fig. 7.3). There was no significant difference between sterilised compost

extracts and the control on sporulation by F. oxysporum (Fig. 7.3). Again, this

result is consistent with a role for living microbiota in the suppressive effect, as

opposed to allelochemicals present in the compost extract.

Figure 7. 3. Effect of compost extracts in trial 1 on sporulation (number of spores per plate) of
Fusarium oxysporum, f. sp. lycopersici six days after inoculation. Cont - control (water), S ACE1
- sterilised aerated compost extract 50 mL L-1, S ACE2 - sterilised aerated compost extract 100
mL L-1, S NCE1 - sterilised non-aerated compost extract 50 mL L-1, S NCE2 - sterilised non-
aerated compost extract 100 mL L-1, ACE1 - aerated compost extract 50 mL L-1, ACE2 - aerated
compost extract 100 mL L-1, NCE1 - non-aerated compost extract 50 mL L-1, NCE2 - non-aerated
compost extract 100 mL L-1. Means with different letters indicate significant difference (p< 0.05)
across treatments and vertical bars represent standard errors of the means (n = 5).

In trial 2, a significant reduction in sporulation by F. oxysporum and F.

solani was achieved by all non-sterilised compost extracts, with no effect of

compost age (three and nine months) or aeration during the extraction process

188
Chapter 7: Compost extract and disease suppression

(Fig. 7.4). This result is consistent with that of Scheuerell (2002), and Scheuerell

and Mahaffee (2000) who noted that both aerated and non-aerated compost

extracts significantly reduced the extent of powdery mildew on rose

(Sphaerotheca pannosa var. rosae), with no difference noted between their

efficacy. No reductions were noted in spore numbers when sterilised compost

extracts (both aerated and non-aerated) were compared to the water control for F.

oxysporum (Fig. 7.4). However, increased numbers of spores of F. solani were

found when treated with nine month old sterilised aerated compost extract (Fig.

7.4). In like manner, sporangia production of the tested fungi (Phytopthora spp.)

was reduced by non-sterilised extracts based on composted Eucalyptus-bark,

while the sterilised extracts stimulated its production (Hardy & Sivasithamparam,

1991).

Figure 7. 4. Effect of compost extracts in trial 2 on sporulation (number of spores per plate) of
Fusarium oxysporum, f. sp. lycopersici and F. solani DAR 66102, Cont - control (water), 3M
ACE - three month old aerated compost extract, 9M ACE - nine month old aerated compost
extract, 9M NCE - nine month old non-aerated compost extract, 9M SACE - nine month old
sterilised aerated compost extract, 9M SNCE - nine month old sterilised non-aerated compost
extract. Means with different letters indicate significant difference (p< 0.05) across a given
species and vertical bars represent standard errors of the means (n = 4).

189
Chapter 7: Compost extract and disease suppression

3.1.3. Total fungal biomass

In trial 1, no treatments reduced the total dry weight of F. oxysporum

compared to the control (Fig. 7.5). Rather, sterilised aerated compost extract

increased the total fungal biomass as compared to that of the water control (Fig.

7.5), possibly because of the loss of antagonistic effect due to sterilisation of

compost extracts on the tested pathogen, as reported by McQuilken et al. (1994).

The sterilised extract might have supplied nutrients or growth promoting

hormones, which presumably enhanced the growth of the tested fungus.

Figure 7. 5. Effect of compost extracts in trial 1 on total biomass (d wt) of Fusarium oxysporum,
f. sp. lycopersici six days after inoculation. Cont - control (water), S ACE1 - sterilised aerated
compost extract 50 mL L-1, S ACE2 - sterilised aerated compost extract 100 mL L-1, S NCE1 -
sterilised non-aerated compost extract 50 mL L-1, S NCE2 - sterilised non-aerated compost extract
100 mL L-1, ACE1 - aerated compost extract 50 mL L-1, ACE2 - aerated compost extract 100 mL
L-1, NCE1 - non-aerated compost extract 50 mL L-1, NCE2 - non-aerated compost extract 100 mL
L-1. Means with different letters indicate significant difference (p< 0.05) across treatments and
vertical bars represent standard errors of the means (n = 5).

In trial 2, although three and nine month old compost extracts showed

similar effects in reducing the total biomass of F. oxysporum, F. solani and R.

solani, three month old aerated compost extract was the most effective in

reducing the total dry weight biomass of the former two pathogens (Fig. 7.6). In

190
Chapter 7: Compost extract and disease suppression

contrast, Winterscheidt et al. (1990), cited in Weltzien (1991), reported that non-

aerated extract based on six month aged horse manure compost was more

effective than that aged one year in suppressing downy mildew of cucumber.

Another study reported reduced efficacy of non-aerated compost extract based on

18 to 24 month aged cattle manure and straw compost, compared to that based on

12 month old compost, in inhibiting the germination of Venturia inequalis

(Andrew, 1993).

The non-aerated sterilised compost extracts (nine month old compost)

stimulated the growth of F. solani and R. solani, with total biomass higher by

44% and 84%, respectively, compared to the control treatment (Fig. 7.6). This

result is consistent with the observed increase in linear mycelial growth and spore

counts associated with the sterilised extract treatment. Other studies have also

reported the stimulation of fungal growth with the application of autoclaved or

filter sterilised compost extracts (e.g. Hardy & Sivasithamparam, 1991;

McQuilken et al., 1994). However, no difference was noted between the effects

of non-aerated compost extract and sterilised aerated compost extract in terms of

the total biomass of the tested pathogens (Fig. 7.6).

Of interest, the non-aerated compost extract (nine month old) did not

reduce the biomass of F. oxysporum and F. solani compared to the control; rather

the total biomass of R. solani treated with nine month old non-aerated compost

extract showed an increased biomass compared to the control (Fig. 7.6). This

clearly showed that non-aerated compost extracts did not combat the fungal

disease organism R. solani, whereas the aerated extracts did. In contrast, Ma et al.

(2001b) showed a mycolytic effect of non-aerated compost extracts on

chlamydospores and microspores of F. oxysporum f. sp. vasinfectum.

191
Chapter 7: Compost extract and disease suppression

Scheuerell and Mahaffee (2002) summarized a number of studies,

although not peer reviewed, conducted in controlled conditions on aerated

compost extracts and found considerable variation in the efficacy of aerated

extracts against various plant diseases.

Figure 7. 6. Effect of compost extracts in trial 2 on total biomass (d wt) of Fusarium oxysporum,
f. sp. lycopersici (black bar), F. solani DAR 66102 (light grey bar), and Rhizoctonia solani DAR
61830 (dark grey bar), six days after inoculation. Cont - control (water), 3M ACE - three month
old aerated compost extract, 9M ACE - nine month old aerated compost extract, 9M NCE - nine
month old non-aerated compost extract, 9M SACE - nine month old sterilised aerated compost
extract, 9M SNCE - nine month old sterilised non-aerated compost extract. Means with different
letters indicate significant difference (p< 0.05) across a given species and vertical bars represent
standard errors of the means (n = 4).

3.2. Hydroponics trials

In trial 3, compost extracts (both sterilised and non-sterilised) were

effective in reducing the disease severity caused by F. oxysporum of tomato

plants grown in the hydroponics compared to the control treatment (Fig. 7.7).

There was no significant difference (p=0.63) in disease severity between tomato

plants treated with either sterilised or non-sterilised compost extract although

there was a clear difference in the performance of either extract on mycelial

growth, spore count and fungal biomass in vitro. Several studies have

192
Chapter 7: Compost extract and disease suppression

demonstrated that sterilisation or micro-filtration does not necessarily affect the

suppressive activity of compost extracts against various phytopathogens (Cronin

et al., 1996; Elad & Steinberg, 1994; Yohalem et al., 1994). Presumably sterilised

compost extract contains an antibiotic that acts to suppress the disease organism,

as discussed by Brinton (1995) and Scheuerell and Mahaffee (2002), or acts to

induce a systemic acquired resistance within the treated plants (Cook et al., 1995;

Litterick et al., 2004; Zhang et al., 1998). The live compost extracts could also

inhibit pathogen growth through parasitism or competition.

Figure 7. 7. (A) A graph showing an effect of application of compost extracts in hydroponics


trial 3 on two month old tomato plants against the severity of fungal wilt caused by F. oxysporum
f. sp. lycopersici grown in the greenhouse (B) pictures of pathogen inoculated control plants (on
the left) and plants treated with compost extracts (on the right). Disease severity based on rating
scale 0-5, 0, no wilt, healthy plant; 1, slightly stunted, <5 % leaves affected; 2, slight yellowing
and wilting, <25 % leaves affected; 3, moderate wilting, 25-50% leaves affected; 4, severe
wilting, 50-75% leaves affected; and 5, >75 % leaves wilted. Compost extracts were applied to
tomato roots until runoff. Cont - control, ACE - Non-sterilised aerated compost extract, S ACE -
Sterilised aerated compost extract (n = 3). Means with different letters indicate significant
difference (p< 0.05) across treatments and vertical bars represent standard errors of the mean (n =
3).

193
Chapter 7: Compost extract and disease suppression

The presence of F. oxysporum was confirmed through axenic culture of

root segments. Roots of tomato plants not treated with compost extract were

found to be significantly infected with F. oxysporum, compared to roots treated

with either non-sterilised or sterilised aerated compost extracts (p< 0.001). Roots

of aerated compost extract treated tomato, however, were assessed to have a low

level of infection, possibly indicating cross contamination (Table 7.1).

Table 7. 1. Fusarium infection in roots based on rating scale of 0 (pathogen absent) to 1


(pathogen present) on eight tomato plant roots in each hydroponics container (trial 3). Cont -
control, ACE - Non-sterilised aerated compost extract, S ACE - Sterilised aerated compost
extract. Means with different letters within rows indicate significant difference (p< 0.05) across
pathogen inoculated and non-inoculated treatments (n = 3).
Treatments Cont ACE S ACE
Pathogen inoculated 0.54 ± 0.10a 0.25 ± 0.12b 0.08 ± 0.08b
Pathogen non-inoculated 0.00 ± 0.00a 0.04 ± 0.04a 0.00 ± 0.00a

In trial 4, across all three pathogens, there was an effect of all compost

extracts in maintaining the level of severity below that of the water control. This

was evident by the third week of the trial but only reached significance in the

fifth week, and then only for F. oxysporum and for nine month old aerated

compost extracts (Fig. 7.8). Earlier in vitro studies (trials 1 and 2) had shown

disease suppression with the use of compost extracts when compared to the

control as revealed by mycelial growth, spore counts as well as total biomass of

all fungal pathogens. Significant inhibition of mycelial growth (up to 70%) was

also shown by Hibar et al. (2006) when the efficacy of compost extract was tested

in vitro on F. oxysporum f. sp. radicis-lycopersici.

No difference was demonstrated between aerated and non-aerated extracts

or between sterilised and non-sterilised extracts in terms of reducing disease

194
Chapter 7: Compost extract and disease suppression

severity (Fig. 7.8). Evidently, the beneficial effect of compost extract was not

clear because severity response of tomato plants varied with different fungal

pathogens despite the strong effect of extracts against all the tested pathogens in

vitro. Hibar et al. (2006) demonstrated an inhibitory role of compost extract

against fusarium crown and root rot disease of tomato crop.

195
Chapter 7: Compost extract and disease suppression

A 4.5 Cont
3M ACE
4.0 9M ACE
9M NCE
9M SACE
3.5 9M SNCE

3.0

Severity (%)
2.5

2.0

1.5

1.0

0.5

0.0
1 2 3 4 5

B 4.5 Cont
3M ACE
4.0 9M ACE
9M NCE
9M SACE
3.5 9M SNCE

3.0
Severity (%)

2.5

2.0

1.5

1.0

0.5

0.0
1 2 3 4 5

C 4.5 Cont
3M ACE
4.0 9M ACE
9M NCE
9M SACE
3.5 9M SNCE

3.0
Severity (%)

2.5

2.0

1.5

1.0

0.5

0.0
1 2 3 4 5
Week after treatment

Figure 7. 8. Effect of application of compost extracts in hydroponics trial 4 on two month old
tomato plants against the severity of fungal wilt diseases caused by (A) F. oxysporum, (B) F.
solani and (C) R. solani in greenhouse assays. Disease severity was based on a rating scale of 0 –
5: 0, no wilt, healthy plant; 1, slightly stunted, <5% leaves affected; 2, slight yellowing and
wilting, <25% leaves affected; 3, moderate wilting, 25 – 50% leaves affected; 4, severe wilting,
50 – 75% leaves affected; and 5, >75% leaves wilted. Compost extracts were applied to tomato
roots until runoff. Cont - control (water), 3M ACE - three month old aerated compost extract, 9M
ACE - nine month old aerated compost extract, 9M NCE - nine month old non-aerated compost
extract, 9M SACE - nine month old sterilised aerated compost extract, 9M SNCE - nine month
old sterilised non-aerated compost extract. Values are mean ± SE (n = 3).

196
Chapter 7: Compost extract and disease suppression

In trial 4, as expected, tomato plants responded well to all of the compost

extracts in terms of chlorophyll concentration, starting from three weeks after

treatment, with significantly higher values in comparison to those of the control

treatment which decreased with time (Fig. 7.9). It was not until the third week

after treatment application that tomato plants inoculated with F. oxysporum and

R. solani showed any greater chlorophyll than in the control (Fig. 7.9A and C).

Four weeks after treatment application, chlorophyll concentration was

significantly higher in the leaves of nine month old aerated compost extract

treated plants which was inoculated with R. solani compared to the inoculated

control (Fig. 7.9). Thereafter, all of the F. oxysporum inoculated plants responded

positively to compost extract application when compared to the water-control

plants (Fig. 7.9). This could simply be attributed to the increased uptake of

nutrients needed for chlorophyll development or to the bio-efficiency of compost

extracts as reported by Siddiqui et al. (2008b), who studied the incidence of wet

rot in okra using Trichoderma-fortified compost extracts. Such a bio-efficiency

could also be attributed to the production of antibiotic compounds, such as

siderophores (Brinton et al., 1996; Potera, 1994) or to induced systemic resistance

developed by tomato plants (Cook et al., 1995; Zhang et al., 1998).

197
Chapter 7: Compost extract and disease suppression

A 40
Cont
9M ACE

Chlorophyll concentration per plant (SPAD units)


9M NCE
9M SACE
35 9M SNCE
3M ACE
30

25

20

15

10

0
1 2 3 4 5

B 45
Cont
9M ACE
Chlorophyll concentration per plant (SPAD units)

9M NCE
40 9M SACE
9M SNCE
3M ACE
35

30

25

20

15

10

0
1 2 3 4 5

C 45 Cont
9M ACE
Chlorophyll concentration per plant (SPAD units)

9M NCE
40 9M SACE
9M SNCE
3M ACE
35

30

25

20

15

10

0
1 2 3 4 5
Week after treatment

Figure 7. 9. Effect of application of compost extracts in hydroponics trial 4 on the leaf


chlorophyll concentration of two month old tomato plants grown in the greenhouse and inoculated
with three different fungal pathogens, namely, (A) F. oxysporum, (B) F. solani and (C) R. solani,
measured one week after the second inoculation. Cont - control (water), 3M ACE - three month
old aerated compost extract, 9M ACE - nine month old aerated compost extract, 9M NCE - nine
month old non-aerated compost extract, 9M SACE - nine month old sterilised aerated compost
extract, 9M SNCE - nine month old sterilised non-aerated compost extract. Values are mean ± SE
(n = 3).

198
Chapter 7: Compost extract and disease suppression

4. Conclusions

In vitro studies show that non-sterilised compost extracts have the ability

to suppress fungal wilt of tomatoes, mostly due to the biotic components of

compost extracts while sterilised extracts in the main enhanced the in vitro

growth of the three wilt pathogens. These results suggest that microbes present in

compost extracts are inhibitory and possibly produce a biochemical that reduces

linear growth of mycelia and inhibits conidia production by the fungal pathogens.

Results of hydroponics trials in vivo showed that inhibition by compost extract

occurs directly or indirectly through nutrient supplement and subsequent

strengthening of the plants on the host-pathogen system. However, the variable

results from the hydroponics trials show that all fungal diseases cannot be

suppressed absolutely by reliance upon compost extract alone. Further

experiments are necessary to identify the mechanisms responsible for the

suppression of diseases observed in the current study.

199
Chapter 8: Compost extract and aluminium toxicity

Chapter 8

The effect of compost extract on root elongation of

maize (Zea mays L.) in the presence of aluminium

Abstract

The aim of this study was to assess the ameliorative effect of compost

extract, used either in solution culture or in seed imbibitions, and either ‗fresh‘ or

‗sterilised‘, on the root growth of maize (Zea mays L. cv. Hycorn 675 IT)

seedlings. Root elongation of maize was inhibited to 43% of the control in the

presence of 200 µM of trivalent aluminium at pH 4.5. Organic acids can chelate

aluminium ions, but the inhibition was not alleviated by either the addition of a

1:500 dilution of microbially enhanced extract of a compost based on rumen

waste or of leachate from vermicasts produced from the same material

(containing 1.6 and 4.4 mg L-1 humic acid, and 2.0 and 1.2 mg L-1 fulvic acid,

respectively). However, when seeds were imbibed in the neat extracts, and then

germinated in the absence of further extract, root elongation was significantly

enhanced in seeds that were imbibed with a fresh compost extract, but not by the

sterilised extract or the vermicast leachate compared to those imbibed in water

(control) and germinated in the same condition. It could be that compost extract

contains an auxin or gibberellic acid-like growth regulator (that is, destroyed by

200
Chapter 8: Compost extract and aluminium toxicity

heat in a sterilised extract), and this could have improved root elongation or

increased germination rate of maize.

1. Introduction

Acidification of agricultural lands threatens sustainable production of

crops and pastures, primarily through increased bio-availability of phytotoxic

aluminium in the soil solution (Clune & Copeland, 1999). The problem becomes

acute at soil pH below 5.0 (Foy, 1974; Foy et al., 1978). Aluminium is available

in multiple forms (e.g. dissolved, bound, free, monomeric, polymeric) in the soil,

and pH can determine its speciation (Imray et al., 1995). The freely available

form of trivalent aluminium (Al+++), largely restricted to acidic conditions, is

phytotoxic (Delhaize & Ryan, 1995).

Nosko et al. (1988) reported that aluminium does not inhibit seed

germination but it inhibits growth of roots after the seeds have germinated,

primarily through an effect on the root apex (Bennet & Breen, 1991; Ryan et al.,

1993). Toxicity symptoms are usually visible as ‗stunted‘ root tips and

‗thickened‘ lateral roots, later turning brown in colour (Roy et al., 1988).

Aluminium ions bind to the DNA of root meristematic cells, hindering mitotic

cell division and root elongation (Clarkson, 1969; Matsumoto et al., 1976;

Matsumoto et al., 1977). Thus, root growth is more sensitive to soil aluminium

than is shoot growth (Taylor, 1988). Affected roots are inefficient in absorbing

nutrients and water from the soil (Foy, 1992), and hence they will become

sensitive to drought and nutrient deficiency. The failure in efficient uptake of

nutrients, particularly of Ca2+ and P, is the secondary physiological effect

associated with aluminium toxicity (Henning, 1975).

201
Chapter 8: Compost extract and aluminium toxicity

In general, the acid-aluminium stress also negatively impacts the growth

and multiplication of microorganisms, for example, Rhizobium (Keyser &

Munns, 1978). Nearly dividing cells show sensitivity to toxic aluminium, while

non-dividing cells can survive an acid-aluminium stress (Munns & Keyser,

1981). Aluminium tolerance has been reported to vary widely between microbial

strains, but remains stable within a given strain (Munns, 1977).

Plant roots are reported to release organic acids in the soil in response to a

range of environmental stresses (Ma et al., 2001). Amelioration of aluminium

toxicity in both soil and solution culture has been achieved through chelation of

the aluminium ions by organic acids, for example, by fulvic acid (65 mg L-1 of

organic C per 3.5 L of nutrient solution), as evidenced by tap root elongation of

seedlings of soybean, cowpea, and green gram grown in solution culture treated

with 50 µM aluminium at pH 4.5 (Suthipradit et al., 1990). Tan and Binger

(1986) conducted a greenhouse experiment to test the effect of humic acid on

corn seedlings grown in sand culture in the presence of 0, 390 and 820 µM

aluminium. In the absence of humic acid, total dry weight of corn seedlings was

reduced by 41% at 820 µM aluminium compared to the control treatment.

However, addition of humic acid at 100 mg kg-1 and 350 mg kg-1 sand increased

the total dry weight of treated corn plants (820 µM aluminium) by 32.5% and

42.5%, respectively, compared to the treatment that had no humic acid.

Harper et al. (1995) tested the effects of 40, 120, 360 mg C L-1 humic and

fulvic acids, extracted from the leaves of Eucalyptus camaldulensis, on root

elongation of Zea mays in solution culture at pH 4.5 with 30 µM aluminium. The

negative effect of aluminium on root elongation was nullified by the addition of

any concentration of humic and fulvic acids. The studies of Harper et al. (1995)

202
Chapter 8: Compost extract and aluminium toxicity

involved solution culture and employed ‗pure‘ humates. Hue and Amien (1989)

reported a soil (pot culture; soil-water pH of 4.0) experiment which demonstrated

the efficacy of different green manures (ground leaves of cowpea, leucaena and

guinea grass) in alleviating aluminium toxicity. They found that aluminium was

detoxified more efficiently by cowpea and leucaena than by guinea grass as

indexed by the biomass of a test crop (Sesbania cochinchinensis) and soil

chemical composition. They concluded that an increment in soil pH with the

addition of manure and chelating action of organic molecules could be the

possible mechanism for aluminium detoxification.

Root elongation is also reported to be stimulated by humic and fulvic

acids at concentrations ranging from 25 to 250 mg C L-1 in the absence of

aluminium (Linehan, 1976; Rauthan & Schnitzer, 1981; Schnitzer & Poapst,

1967; Vaughan & Linehan, 1976). However, Harper et al. (1995) found that

while low concentration (40 mg C L-1) of humic acid stimulated root elongation,

a high concentration of humic acid (360 mg C L-1) reduced root length, both in

the presence or absence of aluminium.

Vaughan and Linehan (1976) suggested that humic and fulvic acids can

have a ‗biochemical or hormonal function‘, while Maggioni et al. (1987)

suggested that these acids can directly affect ATPase activity and influence

nutrient (Mg2+, K+) uptake by plants. Low concentrations of humic and fulvic

acids (0 to 10 mg C L-1) were associated with enhanced in vivo levels of

indoleacetic acid oxidase activity (Mato et al., 1972a; b). However, activity of

this enzyme decreased with higher concentrations of soil humic acid fractions (10

to 40 mg C L-1) (Mato et al., 1972a). They established a direct relationship

between the extent of enzyme activity inhibition and the number of phenolic

203
Chapter 8: Compost extract and aluminium toxicity

groups, ratio of C to H, and the C and N concentrations of humic acids. Muscolo

et al. (1999) reported a stimulation of nitrogen metabolism and cell growth in

Daucus carota by humates. Others (Aibuzio et al., 1986; Piccolo et al., 1992)

found that nutrient uptake and plant growth regulation are triggered by humic

substances.

Compost extracts (‗compost teas‘) are in relatively common use in a range

of production systems (from biodynamic through organic to broadacre grain

cropping, see Chapter 1), primarily for their perceived role in enhancing soil

microbial activity (i.e. ‗soil health‘). However, such extracts appear to be a rich

source of organic acids, such as humates and fulvates in rumen compost extract

(Shrestha et al., 2011a), as measured and judged from their visual deep dark

coloured appearance, and thus may play a role in alleviation of aluminium

toxicity in acid soils. Leachates from earthworm processed composts appear even

darker in colour than extracts of the raw compost. Therefore, experiments were

undertaken to evaluate the effectiveness of compost and vermicast extracts in

alleviating aluminium toxicity, in the context of the presence of humic and fulvic

acids in the extracts. The measured levels of humates (0.7 g L-1) and fulvates (1.1

g L-1) (Chapter 3, also reported as Shrestha et al., 2011a) are one to three orders

of magnitude lower than those reported in studies employing addition of ‗pure‘

humates.

Compost extracts (e.g. vermicast leachates) contain a high microbial load,

as well as organic acids (Shrestha et al., 2011a). It has been suggested that

earthworms promote microbial activity in compost producing significant amounts

of plant growth hormones such as indole-acetic acids (auxins), gibberellins and

cytokinins (Edwards, 1998). Casenave de Sanfilippo et al. (1990) also found

204
Chapter 8: Compost extract and aluminium toxicity

contents of such plant growth hormones in humic acids produced from

earthworm activities. It is possible that the active microbiota of compost extracts

may directly take up aluminium ions, thus reducing the toxic aluminium

concentrations in the root zone, or compost extract may consist of plant growth

regulators that enhance root growth.

The current study assesses the ameliorative effects of compost extract,

used either in solution culture or in seed imbibitions, and either ‗fresh‘ or

‗sterilised‘, on the growth of maize roots.

2. Materials and methods

2.1. Rumen compost extract and vermicast leachate

A microbially enhanced extract of rumen content (cattle paunch) compost

was produced as described in Chapter 3 (briefly, a 1:10 ratio of compost to water

was amended with 1% ‗fish and kelp hydrolysate‘ and 0.5% molasses and

incubated with aeration for 24 h to produce rumen compost extract, ‗RCE‘). The

vermicast leachate, ‗VL‘ used in this experiment was obtained from feeding the

above mentioned compost to a mixture of three species of earthworms (as

described in Chapter 3). The umen compost extract and vermicast leachate were

used ‗fresh‘ or ‗sterilised‘ (121˚C for 15 min at 0.1034 MPa). The extracts were

filtered through 0.45 µm-pore membrane filters (Millipore, Australia) to ensure

elimination of live microbes in the extracts.

2.2. Chemical and microbial analyses of compost extracts

Inorganic nutrients, humic and fulvic acid concentrations of both sterilised

and non-sterilised rumen compost extract and vermicast leachate collected after

205
Chapter 8: Compost extract and aluminium toxicity

24 h of incubation were analysed following the methods as described in Chapter

2.

Culturable populations of bacteria and fungi of serially diluted samples

were determined by the pour plate count method (Harrower, 2006) as described

in Chapter 2.

2.3. Plant material and experimental set up

Root weight is reported to be less sensitive to aluminium damage than

root length (Jackson, 1967). Thus, only tap root elongation was considered in the

present study. The bioassay procedures used are based on those of Clune and

Copeland (1999) and Harper et al. (1995), with modifications noted below.

Seeds of maize (Zea mays L. cv. Hycorn 675 IT; courtesy of John Eggins,

Pacific Seeds, Australia) were washed in running tap water for 15 minutes and

rinsed with distilled water to remove fungicide. They were then soaked in a

solution of 200 μM CaSO4 and 50 μM H3BO3 for 30 minutes.

The experimental set up consisted of a 1 L plastic tray containing 500 mL

of reverse osmosis (RO) water, upon which a styrofoam block (13 cm long x 9

cm wide x 1.5 cm thick) was floated. The styrofoam was covered with autoclaved

paper towels (20.2 x 28.5 cm2) (Kimsoft Kleenex brand, Australia), and five

seeds of maize were placed onto the raft. Solution pH was adjusted to and

maintained at 4.5 by addition of 0.1 M HCl or NaOH. Seeds were allowed to

germinate and grow on the raft for 96 h, with trays held under ambient laboratory

conditions (approximately 22˚C).

In a preliminary trial, treatments consisted of amendment of the tray water

to achieve 0, 25, 50, 100, 200, 400, 600, 800 and 1000 μM Al (pH 4.5), through

206
Chapter 8: Compost extract and aluminium toxicity

addition of a stock solution of 0.5 M AlCl3. After imposition of the treatments,

seedlings were allowed to grow for a further 96 h and tap root length was then re-

measured, with data expressed as the average root length extension per seed, per

tray. Four replicate trays were maintained for each treatment. This preliminary

experiment was undertaken to establish the concentration of aluminium that

suppressed root extension under the conditions of the bioassay. A single

concentration was then imposed as a treatment relative to a zero aluminium

control in two further experiments.

In the main experiment, seedling root length was measured, and then the

aluminium treatments imposed. Then the seedlings were allowed to grow for a

further 96 h, during which the seedling root extension was recorded.

In the first experiment, the compost extracts (rumen compost extract and

vermicast leachate, both non-sterilised and sterilised) were added to the solution

culture at 1 mL per 500 mL of aluminium contaminated solution (200 µM Al) or

pure water. In the second experiment, maize seeds were imbibed overnight (12 h)

in the undiluted compost extracts (non-sterilised or sterilised) or water (control),

before placing them on styrofoam blocks contained in the plastic trays. Seedlings

were grown on the rafts floating on water (pH 4.5) for 96 h but without further

addition of compost extract.

Experimental set up was the same for all experiments, including the

preliminary trial, however the treatments differed.

2.4. Statistical analysis

Significant differences among treatments were determined using simple

analysis of variance using XLSTAT (2010) at a confidence level of 95%.

207
Chapter 8: Compost extract and aluminium toxicity

3. Results and discussion

3.1. Effect of Al on root growth of maize seedlings - Preliminary trial

Maize seedlings exposed to > 25 μM aluminium showed a reduction in

root extension, with the effect leveling off at concentrations above 200 μM (at

55% of control; Fig. 8.1). Clune and Copeland (1999) reported a reduction in

relative root extension of canola seedlings when exposed to aluminium (from 40

to 100 μM) at a pH of 4.5 in nutrient solutions as compared to non-treated

control, with a decrease to 55% of control noted at 80 μM Al.

Figure 8. 1. Relative root extension (% of control, where control plants exhibited a root extension
of 14 cm) of maize (Zea mays L. cv. Hycorn 675 IT) grown in a solution culture (pH 4.5) at a
range of aluminium concentrations. Vertical bars represent standard errors of the means (n = 4).

The 200 μM aluminium was chosen as the base treatment in assessing

the effects of compost extract treatments on maize seedling root growth.

208
Chapter 8: Compost extract and aluminium toxicity

3.2. Composition of compost extract

The humic acid concentration of vermicast leachate was twice as that of

rumen compost extract (at approximately 2 mg L-1). In contrast, fulvic acid

concentration of vermicast leachate was lower than that of rumen compost extract

(Table 8.1). Presumably earthworm activity selectively enhances humic acid

content. Bacterial and fungal loads were higher in the rumen compost extract, as

expected for a microbially enhanced product (i.e. food sources and aeration

provided), relative to the vermicast leachate (Table 8.1). Thus, microbial load

also varied between these treatments. Inorganic ion concentrations (NH4+-N,

PO4−-P, and K+-K) were higher in rumen compost extract than in vermicast

leachate (Table 8.1), and hence it is possible that the osmolality of the solutions

could impact on germination rates. Following seed germination, however, root

elongation should be relatively independent of rhizosphere nutrient availability, at

least for the first few days of growth, as the seeds rely upon available seed

reserves.

Table 8. 1. Comparison of physico-chemical and microbial characteristics of sterilised and non-


sterilised rumen compost extract and vermicast leachate. RCE - rumen compost extract, S RCE -
sterilised rumen compost extract, VL - vermicast leachate, S VL - sterilised vermicast leachate.
Within columns, mean ± SE (n = 3) with the same letter are not significantly different according
to Tukey‘s test (p< 0.001).
Parameters RCE S RCE VL S VL
+ -1
NH4 -N (mg L ) 1.03 ± 0.04a 0.98 ± 0.05a 1.45 ± 0.04b 1.42 ± 0.03b
- -1
PO43 -P (mg L ) 21.3 ± 0.78a 20.0 ± 1.52a 27.0 ± 1.21b 27.6 ± 0.78b
+ -1
K -K (g L ) 0.67 ± 0.02b 0.66 ± 0.02b 0.44 ± 0.03a 0.45 ± 0.02a
-1
Humic acid (g L ) 0.77 ± 0.06a 0.79 ± 0.08a 2.15 ± 0.29b 1.95 ± 0.02b
-1
Fulvic acid (g L ) 0.99 ± 0.02b 0.92 ± 0.08b 0.63 ± 0.03a 0.64 ± 0.04a
Bacterial population (cfu Log 10 mL-1) 11.5 ± 0.03c 0.00 ± 0.00a 6.72 ± 0.07b 0.00 ± 0.00a
Fungal population (cfu Log 10 mL-1) 5.29 ± 0.06c 0.00 ± 0.00a 4.20 ± 0.10b 0.00 ± 0.00a

209
Chapter 8: Compost extract and aluminium toxicity

3.3. Experiment 1

The presence of vermicast leachate in the hydroponic solution (1:500

dilution; both ‗fresh‘ and ‗sterilised‘) significantly (p< 0.05) decreased root

extension, compared to the water control (Fig. 8.2). This is particularly

pronounced in the presence of Al. Compost extract did not significantly improve

or retard root extension (Fig. 8.2). Vermicast leachate, which contains substances

that decrease root growth, affected root extension even at a dilution of 1:500.

This response is consistent with that noted by Churilova (2010) in pak choi plants

treated with the same vermicast leachate (1:1, 1:2 dilutions) compared to that of

the control plants.

Figure 8. 2. Experiment 1: Root elongation rate of maize (Zea mays L. cv. Hycorn 675 IT)
seedlings (n = 4) following growth for 96 h on a solution culture with or without 200 µM of Al at
pH 4.5, with treatments of non-sterilised and sterilised rumen compost extracts and vermicast
leachate (1:500 dilution). The control consisted of water only. Cont - control, RCE - rumen
compost extract, S RCE - sterilised rumen compost extract, VL - vermicast leachate, S VL -
sterilised vermicast leachate. Bars with the different letters indicate significant difference (p<
0.05). Error bar represents one standard error of the mean (n =4).

Neither vermicast leachate nor compost extract treatments alleviated the

inhibition of root extension in the presence of 200 μM aluminium, relative to the

210
Chapter 8: Compost extract and aluminium toxicity

control (Fig. 8.2). At a dilution of 1:500, the compost extract treatment would

have contained 1.6 and 2.0 mg L-1, that is, 0.6 and 0.8 mg C L-1 (assuming these

compounds have 40% C), or 0.8 and 1.0 mg tray-1 humic and fulvic acids,

respectively. These levels are low compared to those used in published literature

(40-360 mg C L-1) on root elongation of maize (Harper et al., 1995). Indeed, the

500 mL of 200 μM aluminium solution in each tray contained 100 µmoles of

aluminium. Assuming a chelation capacity or cation exchange capacity of 300

cmol+ kg-1, for example, as for soil organic matter or humus (DPIPWE, 2010),

the humates and fulvates present in each tray could potentially complex 1.8 mg

tray-1 x 0.3 x 0.001 mg g-1 = 5.4 µmoles. Thus, on a stoichiometric basis, it

appears that the applied level of humates was inadequate to complex the available

Al. To achieve stoichiometry, the compost extract should ideally for this

proposed effect have been used at a dilution of at least 1:25 (instead of 1:500).

The level of compost extract addition employed in the current study was

chosen as a level that had practical relevance (in soil, roots will not be exposed to

‗neat‘ compost extracts). These results show that the above extracts do not

ameliorate soil aluminium toxicity. However, humates may accumulate in the soil

from successive additions of compost extracts and this accumulation may lead to

reduced aluminium toxicity. Tejada et al. (2010) compared the effects of various

organic amendments (e.g. municipal solid waste compost, poultry manure and

cow manure) on aluminium contaminated soil using an earthworm incubation

experiment. The authors found that the amendments with higher level of humic

acids were more effective in reducing aluminium toxicity in soil (as indicated by

the earthworm response) compared to those with lower levels. Any future work in

this area should, therefore, attempt to achieve at least an order of magnitude high

211
Chapter 8: Compost extract and aluminium toxicity

level of humic and fulvic acids (e.g. through use of a different compost or

through concentration of the extract), and should consider the issue of soil

accumulation of humates.

3.4. Experiment 2

Pre-imbibed seed in compost extract showed an increased tap root

elongation (Fig. 8.3) compared to that in Experiment 1 (Fig. 8.2). Seed priming is

well known to improve seedling growth (Clark et al., 2001; Rashid et al., 2004).

Presumably the 12 h imbibition treatment of Experiment 2 allowed for quicker

imbibition than that the wet towel-raft conditions of Experiment 1.

Root elongation of maize seeds imbibed in non-sterilised rumen compost

extract (non-diluted) and grown in 200 µM aluminium solution was increased by

51%, compared to the control, whereas that in sterilised compost extract or

vermicast leachate was not significantly different to that of the control treatment

(Fig. 8.3). The increase in root extension with the compost extract treatment

cannot be explained by a ‗fertilisation‘ effect, given the lack of response to the

vermicast leachate treatments.

212
Chapter 8: Compost extract and aluminium toxicity

Figure 8. 3. Experiment 2: Root elongation rate of maize (Zea mays L. cv. Hycorn 675 IT)
seedlings (n = 4) measured 96 h after imbibition in solution culture in the presence or absence of
200 µL Al (solution pH 4.5). Seeds were imbibed for 12 h in non-sterilised and sterilised rumen
compost extracts, vermicast leachate and the water control. Cont - control, RCE - rumen compost
extract, S RCE - sterilised rumen compost extract, VL - vermicast leachate, S VL - sterilised
vermicast leachate. Bars with the different letters indicate significant difference (p< 0.05), and
error bar represents one standard error of the mean (n = 4).

The lack of response of the sterilised compost extract may be ascribed to

either the active role that microorganisms might have in the uptake of aluminium

from the solution, to the loss of chelating capacity of the extract due to

sterilisation, or to the loss of a growth regulator (e.g. auxin) activity.

During imbibition, solution is osmotically drawn into the seed. The

physical flow would result in the movement of solutes contained in the solution

(including humic and fulvic acids) into the seed. However, these chemicals are

unlikely to move within the plant. After 96 h of germination, roots were

approximately 6-8 cm long (i.e. in the presence of Al). The Al sensitive root tips

presumably at this stage, and somewhat earlier, did not contain humic or fulvic

acid from the initial imbibition treatment. Rather any difference between

treatments must be ascribed to a physiological effect of these chemicals on the

germinating seedling, for example, through an increase in gibberellins or auxin

213
Chapter 8: Compost extract and aluminium toxicity

levels, leading to an earlier germination or an increase in the root extension rate

(Mato et al., 1972a; b).

4. Conclusion

Although organic acids like humic and fulvic acids within compost

extract were expected to chelate aluminium ions, the inhibition of root elongation

by trivalent aluminium ions was not improved by either microbially enhanced

extracts of compost based on rumen waste or of leachate from vermicasts

produced from the same material. In contrast, maize seeds imbibed with fresh

compost extract demonstrated significantly enhanced root elongation in the

presence of aluminium ions, whereas seeds imbibed with sterilised extracts or the

vermicast leachate did not. The influence of compost extract, possibly via growth

regulators present in the compost extracts, on rate of germination should

be considered.

214
Chapter 9: Compost extract and soil microbial activity

Chapter 9

Enhancing microbial activity of a field soil with

long-term use of liquid biofertilisers

Abstract

The influence of several agricultural management practices, including

liquid biological inocula, green manures, liquid fertiliser and conventional

chemical fertiliser, on soil physico-chemical and microbial properties of a

cultivated soil at Baralaba in central Queensland, Australia, was investigated with

comparison to soil properties of an adjacent site with the original native

vegetation (brigalow, Acacia harpophylla). The effect of these management

practices was minor compared to that of seasonal changes, and compared to the

difference between the cultivated soil and the brigalow soil. For example, nitrate

levels were 87% higher in brigalow soil than in conventionally cultivated soils

during the dry season. Average soil nitrate level in cultivated soils (average of all

treatments) was 16% higher in the wet compared to the dry season. Microbial

activity was higher in the brigalow soil than soil of any cultivation treatment,

though activity was enhanced by all treatments, including the application of

compost extract, compared to the control. Substrate utilisation patterns

(BiologTM) of bacterial and fungal community functional diversities did not differ

215
Chapter 9: Compost extract and soil microbial activity

between treated field soils and the brigalow soil, however bacterial community

diversity was higher in biologically treated soils compared to that of the

conventionally cultivated soils in the dry-winter season.

1. Introduction

Traditionally, ecosystem health has been assessed in terms of the ‗macro‘

components of the ecosystem, such as land form, vegetation and macrofauna; but

attention is now being given to the diversity of soil microbiota. Generally, soil

quality is characterised by its abiotic factors (e.g. cation exchange capacity/water

holding capacity), however increasing recognition is being given to biotic factors

– the ‗health‘ of the soil (Doran & Safley, 1997; Pankhurst et al., 1997).

Microorganisms are considered to be useful indicators of soil quality due to their

fundamental roles in organic matter decomposition, nutrient recycling and

maintenance of soil structure (Anderson, 2003; Lalande et al., 2000; Robertson et

al., 1994; Sparling, 1997).

Accelerated and intensified use of agricultural soils has negatively

impacted ‗soil health‘ (McCaig et al., 2001; Nusslein & Tiedje, 1999; Ovreas et

al., 1998). Indeed up to 40% of world‘s agricultural land has been estimated to be

seriously degraded (Sample, 2007), with degradation in Australia also reported to

be substantial (Mabbutt, 1992; Woods, 1983). Intensive farming practices, for

example, tillage and biomass removal from fields, result in increased

mineralisation of soil organic matter and thus its depletion (Dalal & Mayer, 1986;

Golchin et al., 1995; Lemenih et al., 2005; Mann, 1986; Oldeman et al., 1990;

Spaccini et al., 2001). Buckley and Schmidt (2001) found lower proportions of

microbial communities (e.g. Proteobacteria, Actinobacteria, as shown by the T-

216
Chapter 9: Compost extract and soil microbial activity

RLFP patterns), in soils subjected to long-term cultivation practices compared to

undisturbed soils. Similarly, Steenwerth et al. (2003) observed variation in soil

microbial community composition associated with cultivation practices, such as

fertiliser application, tillage frequency, grazing, irrigation, and herbicide

application.

Conversely, addition of biological materials (e.g. farm yard manure) has

shown to improve bacterial community composition in a long-term cropping-field

experiment (Widmer et al., 2006). Change in microbial community structure has

also been noted by Peacock et al. (2001), who found that application of dairy

manure increased soil organic C, increased microbial biomass, and enhanced the

community of typical Gram-negative bacteria compared to the control and

ammonium nitrate fertilised treatments. It is generally accepted that increasing

soil organic matter will result in improved soil structure (Pulleman et al., 2003),

and enriched soil biodiversity (Tu et al., 2006).

To restore degraded soils, techniques, such as mulching, crop rotation,

green manuring, use of compost, and more recently, amendment of soils with

liquid biological fertilisers are commonly used (Garcia-Gil et al., 2000;

Hernandez et al., 2007; Omay et al., 1998). The resulting change in soil

microbiota is also considered to convey resistance to soilborne plant pathogens

(Cotxarrera et al., 2002; Zwieten et al., 2007). Microbially enhanced compost

extracts (‗compost tea‘) are one type of liquid biological fertiliser.

The impact of a ‗compost extract‘ is likely to fall into one of two

categories: (i) where the extract is applied in high volume, or in a localised

manner (e.g. with seed), there will be a capacity to increase the overall soil

microbial population in that soil volume; (ii) where the ‗extract‘ is applied at a

217
Chapter 9: Compost extract and soil microbial activity

very low rate, but it contains microbiota that play key roles in various nutrient

cycles, or in protection of plants from pathogens, but are lacking from the soil.

Addition of rhizobia to soil is an example in which the overall soil microbial load

may be unaltered, and overall diversity number not measurably increased, and yet

the performance of a matched legume crop measurably enhanced.

The effect of organic and microbial amendment of soil is expected to be

greater in moist soil. Central Queensland broadacre cropping on black cracking

clays is largely reliant on rainfall, and the soil profile can be effectively emptied

of water by the end of a crop, and through the dry season. In such a system, soil

microbial populations will vary dramatically between seasons, and it is not clear

if a microbial soil amendment alone can have a sustained impact on soil

microbial population.

In 2007, a project (Central Queensland Regenerative Farming Initiative)

was established to conduct a farm based trial (coordinated by Scott Stevens,

funded by the Fitzroy Basin Association ‗FBA‘) of seven management options at

Baralaba on the central Queensland highlands, on a site typical of this cropping

region (cracking clay soils, originally vegetated by brigalow, Acacia

harpophylla). This trial was commenced in 2007, with grain sorghum planted on

3rd September, 2007, followed by wheat, sown on 17th June, 2008, then

mungbean, sown on 19th January, 2009. The trial site management followed the

normal local practice of opportunity cropping under rainfed conditions and zero

tillage, where practical, with cereal (wheat, sorghum) alternated with pulse crops

(mungbean, chickpeas) or green manures (mixture of forages, legumes and

cereals). The current study was designed to ‗value add‘ to the ongoing field study

218
Chapter 9: Compost extract and soil microbial activity

in central Queensland, crop yield assessment work being undertaken, in terms of

a focus on soil microbiological properties.

2. Materials and methods

2.1. Study site and treatments

Field samples were obtained from a farm trial site located at Baralaba,

central Queensland (latitude: 24.24˚ S and longitude: 149.86˚ W). This site has

been continuously cropped for thirty years, following clearing of the original

brigalow (Acacia harpophylla) community. The aim of the trial was to compare

biological farming systems with the current conventional cropping system in

terms of their effect on soil ‗health‘. Seven soil amendment systems were

implemented : (i) ‗best-bet biology‘ (BBB), (ii) ‗biology boom spray‘ (BBS), (iii)

‗biology direct injection‘ (BDI), (iv) feather-top Rhodes (FTR), (v) green manure

(GM), (vi) ‗best-bet conventional‘ (BBC) and the (vii) control (Cont), with

details of each treatment shown in Table 9.1.

These treatment applications were made in 2009 to the mungbean crop

(Table 9.1). The control treatment did not receive any fertiliser. In the best-bet

conventional treatment, recommended rates of fertilisers (Starter Z, Incitec Pivot

Ltd, Vic., Australia) were applied at planting. ‗Starter Z‘ is a zinc-fortified

granular fertiliser consisting of a balance of nitrogen, phosphorus and sulphate

sulphur. In the best-bet biology treatment, compost extract along with additional

nutrients and microbial food was applied via a liquid injection method during

planting, at the rate of 30 L ha-1, while foliar applications of a microbial inoculant

(Twin N diazotrophs, AgriBiotec Pty Ltd, Qld, Mapleton, Australia, which are

marketed as free-living rhizosphere N2 fixers) were made just before onset of

219
Chapter 9: Compost extract and soil microbial activity

flowering at the rates mentioned in Table 9.1. The same soil injection

amendments (compost extract along with additional nutrients and microbial food)

were made for the biology direct injection treatment, but the foliar application

was omitted. In the biology boom spray treatment, the same liquid amendments

as in the best-bet biology and biology direct injection treatments were applied

during planting but the compost extract was omitted, and a foliar (boom spray)

application of microbial inoculant (Twin N) was made prior to the onset of

flowering. The green manure and feather-top Rhodes treated plots were planted

with panorama millet, cowpea, lab lab and forage sorghum as a shotgun mix at

the rate of 4 kg ha-1 each on the same day of other treatment applications. This

was then turned back in, using a disc plough, 7 weeks after planting. The green

manure treatment was specifically set up to look at the effects of introducing

green manures into a biological program. In the feather-top Rhodes treatment,

strategies were adopted to control feather-top Rhodes grass. Both treatments used

similar products (biology and liquid nutrients) to best-bet biology, biology direct

injection and biology boom spray. There was no difference in management

between both these treatments between the sampling dates used to generate the

results within this study. However, prior to soil sampling, these two treatments

differed in that the green manure treatment received Starter Z at planting and

Twin N as a foliar spray while the feather-top Rhodes treatment received calcium

nitrate as a foliar spray for the preceding wheat crop. Rhizobial inoculant

‗Nodulaid‘, from Becker and Underwood (www.beckerunderwood.com), was

applied at the same rate to all treatments. All treatments received equal rates of

herbicide (January, 2009) and insecticide (March, 2009) during the growth of

mungbean crop because the weeds and insect pressure was considered the same

220
Chapter 9: Compost extract and soil microbial activity

across the treatments. The herbicides applied were Haloxyfop 520 (HerbiGuide

Pty Ltd, WA) at 520 g L-1 ha-1, 2, 4 - D MCPA sprays (KensoAgcare, Qld) and 2,

4 - D Amine 625 (Dow AgroSciences Australia Ltd, NSW) at 625 g L-1 ha-1each,

while insecticide, Dimethoate 400 (Conquest Agrochemicals Pty Ltd, WA) was

applied at 0.4 g L-1 ha-1.

Table 9. 1. Details of treatments and fertility programs of a long-term field trial (commenced in
2007) conducted in Baralaba, Qld. Table showing treatment applications made in 2009 to the
mungbean crop.

Treatments Fertility program Application Rate


Control (Cont) Sodium molybdenate (kg ha-1) 0.01
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Best-bet biology (BBB) Compost extract (L ha-1) 30.0
CalSap (L ha-1) 3.00
Liquid N (L ha-1) 2.00
Sodium molybdenate (kg ha-1) 0.05
Fish hydrolysate (L ha-1) 2.00
Humic acid (L ha-1) 0.50
Sea minerals (L ha-1) 1.00
Inoculant (Twin N) (kg ha-1) 0.10
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Biology direct injection (BDI) Compost extract (L ha-1) 30.0
CalSap (L ha-1) 3.00
Liquid N (L ha-1) 2.00
Sodium molybdenate (kg ha-1) 0.05
Fish hydrolysate (L ha-1) 2.00
Humic acid (L ha-1) 0.50
Sea minerals (L ha-1) 1.00
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Best-bet conventional (BBC) Starter Z (kg ha-1) 10.0
Sodium molybdenate (kg ha-1) 0.01
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Biology boom spray (BBS) CalSap (L ha-1) 3.00
Liquid N (L ha-1) 2.00
Sodium molybdenate (kg ha-1) 0.05
Sea minerals (L ha-1) 1.00
Inoculant (Twin N) (kg ha-1) 0.10
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Green manure (GM) Panorama millet seed (kg ha-1) 4
Cowpea seed (kg ha-1) 4
Lab lab seed (kg ha-1) 4
Forage sorghum seed (kg ha-1) 4
CalSap (L ha-1) 3.00
Liquid N (L ha-1) 2.00
Sodium molybdenate (kg ha-1) 0.05
Sea minerals (L ha-1) 1.00
Peat inoculant (Nodulaid) (kg ha-1) 0.10
Feather-top Rhodes (FTR) Panorama millet seed (kg ha-1) 4
Cowpea seed (kg ha-1) 4
Lab lab seed (kg ha-1) 4
Forage sorghum seed (kg ha-1) 4
CalSap (L ha-1) 3.00
Liquid N (L ha-1) 2.00
Sodium molybdenate (kg ha-1) 0.05
Sea minerals (L ha-1) 1.00
Peat inoculant (Nodulaid) (kg ha-1) 0.10

221
Chapter 9: Compost extract and soil microbial activity

The compost extract used in the best-bet biology and biology direct

injection treatments was based on a mixture (1:1:1 ratio) of three different

composts, two from CQ compost Pty Ltd in Emerald (aerobic compost) and the

third (vermicompost) from Nutrigold, Mareeba. The thoroughly mixed compost

(4 kg) was amended with 15 mL each (diluted in 200 mL water) of Aloe-TechTM,

Sea-Change KFFTM liquid fertiliser (Nutri-Tech Solutions products, Yandida,

Qld) and Fish Hydrolysate (South Australian Marine Product Industries, SAMPI,

Fremantle, WA) five days before the extract incubation. The following recipe

was then used for preparing compost extract: 4 kg of compost, 1 L of Sampi Fish,

0.5 L of Aloe-TechTM, and 250 g of oat bran were added to 700 L of rainwater at

the onset of incubation. Compost extract was aerated for 24 h and used within 2

to 4 h following incubation.

Treatments were applied to three replicate plots each, except for the

treatments of green manure and feather-top Rhodes, for which only two replicate

plots were established. Each plot was 600 m by 11.4 m in size. The treatments

were imposed in a randomised complete block design (seven treatments, three

replicate plots).

2.2. Soil sampling

Soil sampling occurred on 19th January, 2009 (wet-summer season) and

on 27th July, 2009 (dry-winter season) (for rainfall data, see Australian Bureau of

Meteorology www.bom.gov.au/climate/glossary/seasons.shtml). Soil was

sampled from each replicate plot. Soil samples (two replications) were also

procured from a nearby intact virgin brigalow community. A 35 cm long

cylindrical stainless-steel corer was used to collect to a depth of 10 cm from the

222
Chapter 9: Compost extract and soil microbial activity

soil surface. Three soil cores were collected from random locations within each

plot. After transport to the laboratory, soil cores from individual plots were well

composited and representative soil samples were taken following the cone

sampling method. The soil samples collected in dry-winter were incubated for

five days after addition of 10% (w/v) water, sieved (2 mm) and stored at 4˚C for

further analysis. No artificial incubation was carried out for samples collected in

the wet-summer.

2.3. Soil characterisation

Soil pH and EC of the samples were measured of 1:5 (w/v) suspensions of

soil in distilled water. Soil moisture and inorganic ion levels were determined

(n=3) as described in Chapters 2 and 3.

Microbial load of triplicate soil samples was assessed using a culture-

based method from serially diluted soil samples according to the method of

Harrower (2006). Microbial activity was determined for each soil sample (i.e.

three analyses per plot) using the fluorescein diacetate hydrolysis method, as

described by Adam and Duncan (2001), and by the basal and substrate induced

respiration (SIR) methods described by Anderson (1982). All of these methods

are described in greater detail in Chapters 2 and 3.

Microbial biomass C was estimated on all soil samples using the

chloroform fumigation extraction method (Vance et al., 1987) (10 g dry weight

each of triplicate soil samples as described in Chapter 3). Total and dissolved

organic C (TOC and DOC) were determined spectrophotometrically at 590 nm

following the dichromate digestion method (Walkley & Black, 1934) and

223
Chapter 9: Compost extract and soil microbial activity

microbial biomass C was calculated following the protocol of Sparling and West

(1988) as described in detail in Chapter 3.

Bacterial and fungal substrate utilisation was assessed of samples

collected in the dry season, using the Ecoplate and FF microplate systems

(BiologTM Inc, USA), respectively. A 100 μL aliquot of a 1 to 10 soil: water

extract was used for each well of the Ecoplate system, while a 150 μL aliquot was

used for the FF microplates. The average well colour development (AWCD) was

calculated at 24 h intervals from 0 h to 168 h for all different C substrates in both

Ecoplate and FF microplate (Konopka et al., 1998), with the 72 h data used for

statistical analysis (Garland, 1996), as described in Chapter 1.

The change in density and diversity of the soil microbial community was

investigated using DGGE, based on bacterial 16S rDNA and fungal ITS regions

amplified from total soil DNA. ‗Total‘ DNA was extracted from 5 g fresh weight

of composite soil samples (i.e. one sample per plot) of control and best-bet

biology treatments in wet-summer and dry-winter seasons (all treatments) using

a soil commercial DNA isolation kit (PowerMaxTM Soil) following the

manufacturer‘s instruction (Mo Bio Laboratories, Inc., Carlsbad, CA), and using

a bead beating method. The extracted DNA from each sample was PCR-

amplified and DGGE analysis was performed for bacterial and fungal

communities (Girvan et al., 2003).

2.4. Statistical analyses

Differences between means were compared for different physical,

chemical and microbial parameters as well as indices of species richness,

evenness and diversity at p< 0.05 using analysis of variance (ANOVA) of the

224
Chapter 9: Compost extract and soil microbial activity

statistical software package (XLSTAT, 2010). The Shannon-Weaver diversity

index H‘ (Shannon & Weaver, 1963) and the equitability index J (Begon et al.,

1990) were also calculated (Dilly et al., 2004; Krebs, 1999) and subjected to

ANOVA.

3. Results

3.1. Physico-chemical, microbial and biochemical analyses

Variations in physico-chemical, biochemical and biological characteristics

were noted among the treatments, with dramatic differences observed between

the two seasons (Table 9.2). Soil pH was not affected by season, but increased by

treatments. EC was only weakly increased by treatment, particularly by green

manure. Moisture content was not different, even between seasons, because

moisture content of soil collected in dry-winter season was measured after adding

water. Nitrate levels were higher in brigalow than the control during the dry

season, but average nitrate level in cultivated soils was 16% higher in the wet

compared to the dry season. Biology direct injection and green manure treatments

were higher in ammonium than in the control, while best-bet conventional

treatment was higher in phosphate compared to other treatments in the summer

season.

In soils of the control treatment, plate counts of bacterial and fungal

populations were higher in the wet-summer than in the (5 day incubated) dry-

winter samples (Table 9.2). Bacterial counts (cfu Log 10) were significantly

higher in best-bet biology, best-bet conventional, biology boom spray and green

manure treatments than in the control in the wet-summer, while counts were

higher in best-bet conventional and biology boom spray treatments, and the

225
Chapter 9: Compost extract and soil microbial activity

brigalow soil, than the control in the (5 day incubated) dry-winter samples (Table

9.1). Fungal counts (cfu Log 10) did not differ significantly between the

treatments in either season; but were significantly higher in the brigalow

community soil. The average microbial biomass C of the seven treatments

(excluding brigalow community) was 47% higher in the wet-summer than in the

dry-winter, as expected given increased temperatures and plant biomass.

However, there was no significant difference in microbial biomass C between

treatments in either season, although the level of the brigalow soil was almost

double than that of the cropping treatments. The attributes of soil microbial

activity, as indicated by FDA, basal and substrate induced respiration, were all

higher in the incubated dry-winter season samples than in the field samples of the

previous wet-summer season (Table 9.2).

In the wet-summer season, basal respiration was significantly higher in

best-bet conventional and green manure treated soils, compared to feather-top

Rhodes, whereas substrate induced respiration was significantly higher in biology

direct injection and feather-top Rhodes soils compared to the control (Table 9.2).

The total microbial activity, however, was the highest in best-bet conventional

treatment, with least microbial activity in green manure during the wet-summer

season (Table 9.2). There were no differences among treatments in terms of total

microbial activity, or basal and substrate induced respiration of soil in the dry-

winter season.

The soil from the brigalow community differed from the other treatments

by virtue of its significantly elevated pH, nitrate, bacterial and fungal populations,

microbial biomass C, substrate induced respiration and total microbial activity in

226
Chapter 9: Compost extract and soil microbial activity

the dry-winter season (Table 9.2). The brigalow forest soil was not sampled in the

wet-summer season.

227
Chapter 9: Compost extract and soil microbial activity

Table 9. 2. Variation in the physico-chemical, microbial and biochemical characteristics of soil samples collected from plots treated with different management practices at
the Baralaba trial site in two different seasons. Cont - control, BBB - best-bet biology, BBC - best-bet conventional, BDI - biology direct injection, BBS - biology boom
spray, FTR - feather-top Rhodes , GM - green manure, BRI - brigalow soil. Values are mean ± SE (n = 3) for all treatments except for green manuring, biological weed
control and brigalow soil samples (n = 2) collected in the dry season. Within rows, within a season, means with the same letter are not significantly different according to
Tukey‘s test (p< 0.05). Moisture content was assessed after addition of 10% (w/v) water in the dry season. ‗-‘ indicates missing values lower than the detectable limit of the
RQflex meter.

Wet season (Summer) Dry season (Winter)


Biochemical and microbial characteristics Cont BBB BBC BDI BBS FTR GM Cont BBB BBC BDI BBS FTR GM BRI
pH (1:5) 5.59 ± 0.0a 7.02 ± 0.0d 6.64 ± 0.0c 7.08 ± 0.1d 7.02 ± 0.0d 6.58 ± 0.0c 6.28 ± 0.1b 6.02 ± 0.1a 6.04 ± 0.1a 6.39 ± 0.0b 6.47 ± 0.0bc 6.72 ± 0.0cd
6.93 ± 0.0d 6.97 ± 0.0d 6.91 ± 0.1d
-1
EC (dS m ) 0.11 ± 0.0a 0.13 ± 0.0ab 0.11 ± 0.0a 0.14 ± 0.0bc 0.14 ± 0.0bc 0.12 ± 0.0ab 0.15 ± 0.0c 0.08 ± 0.0a 0.30 ± 0.0cd 0.08 ± 0.0a 0.35 ± 0.0d 0.25 ± 0.0bcd
0.10 ± 0.0ab 0.11 ± 0.0ab 0.16 ± 0.1abc
Moisture content (%) 13.0 ± 0.1a 14.2 ± 0.0a 13.8 ± 0.4a 13.3 ± 0.2a 13.2 ± 0.5a 12.8 ± 0.1a 14.1 ± 0.4a 13.4 ± 0.2a 13.4 ± 0.2a 13.5 ± 0.1a 13.8 ± 0.3a 14.7 ± 0.5a
14.0 ± 0.2a 13.3 ± 0.5a 19.6 ± 1.3b
- -1 7.79 ± 1.3a 33.3 ± 2.4ab 7.79 ± 0.7a 61.4 ± 27.5b
NO3 -N (mg kg d wt) 26.0 ± 2.0a 26.7 ± 1.6a 21.4 ± 3.6a 27.3 ± 2.0a 23.4 ± 2.0a 36.3 ± 4.7a 29.3 ± 4.6a 29.3 ± 3.2ab 30.6 ± 4.1ab
19.0 ± 8.5ab 32.4 ± 6.5ab
+ -1
NH4 -N (mg kg d wt) 0.45 ± 0.0a 1.35 ± 0.3a 1.80 ± 0.5a 4.91 ± 0.5b 0.59 ± 0.1a 1.78 ± 0.3a 3.61 ± 0.3b 3.87 ± 0.3c 1.19 ± 0.3a 2.68 ± 0.5abc 1.80 ± 0.5abc 1.52 ± 0.3ab
3.61 ± 0.9bc 1.78 ± 0.0abc 2.43 ± 0.5abc
-1
PO43- -P (mg kg d wt) 10.0 ± 1.3a 6.32 ± 0.7a 20.5 ± 2.0b 8.00 ± 0.9a 5.00 ± 0.2a 6.00 ± 0.0a 6.57 ± 0.7a 7.25 ± 1.1a 4.25 ± 0.2a 4.50 ± 0.4a - - 9.48 ± 1.9a 4.50 ± 0.8a 8.15 ± 3.3a
Bacterial population (cfu Log 10) 6.00 ± 0.0a 6.62 ± 0.1c 6.63 ± 0.0c 6.36 ± 0.1b 6.69 ± 0.1c 6.30 ± 0.0b 6.52 ± 0.0bc 5.30 ± 0.2a 5.56 ± 0.0a 6.61 ± 0.0b 5.49 ± 0.1a 6.24 ± 0.1b 5.59 ± 0.1a 5.47 ± 0.2a 6.29 ± 0.0b
Fungal population (cfu Log 10) 4.20 ± 0.1a 4.36 ± 0.1a 4.36 ± 0.1a 4.10 ± 0.1a 4.00 ± 0.0a 4.16 ± 0.2a 4.42 ± 0.1a 3.52 ± 0.0a 3.33 ± 0.2a 3.10 ± 0.1a 3.30 ± 0.2a 3.42 ± 0.1a 3.16 ± 0.2a 3.16 ± 0.2a 4.27 ± 0.1b
-1
Microbial biomass C (mg C g d wt) 0.70 ± 0.4a 1.18 ± 0.3a 1.06 ± 0.5a 0.79 ± 0.1a 0.97 ± 0.5a 0.86 ± 0.2a 1.09 ± 0.2a 0.33 ± 0.0a 0.70 ± 0.3a 0.45 ± 0.2a 0.59 ± 0.1a 0.76 ± 0.2a 0.44 ± 0.0a 0.28 ± 0.1a 1.94 ± 0.3b
-1 -1
Basal respiration (µg CO2 g d wt h ) 16.0 ± 1.8ab 21.6 ± 3.6ab 30.3 ± 3.1b 21.3 ± 4.7ab 17.8 ± 3.5ab 10.7 ± 1.8a 26.0 ± 0.9b 14.1 ± 4.6a 31.0 ± 5.5a 33.6 ± 19a 31.9 ± 7.1a 32.2 ± 1.8a 39.1 ± 1.9a 30.8 ± 2.2a 44.7 ± 7.7a
-1 -1
Substrate induced respiration (µg CO2 g d wt h ) 17.6 ± 4.6a 27.6 ± 2.4ab 33.7 ± 4.7ab 38.8 ± 4.7b 35.2 ± 1.8ab 40.3 ± 4.6b 31.1 ± 0.9ab 24.7 ± 3.0a 33.5 ± 7.7a 35.3 ± 11a 47.9 ± 1.8ab 39.4 ± 1.8ab 35.5 ± 6.2ab 37.9 ± 19ab 78.5 ± 13b
-1 -1
Total Microbial activity (µg FDA g d wt h ) 114 ± 6.3ab 131 ± 2.9ab 150 ± 7.8b 111 ± 4.7ab 119 ± 3.2ab 115 ± 12ab 106 ± 15a 224 ± 13a 254 ± 23a 216 ± 11a 311 ± 13a 275 ± 28a 237 ± 58a 269 ± 116a 508 ± 110b
d wt Dry weight

228
Chapter 9: Compost extract and soil microbial activity

3.2. Community level physiological profiles of bacteria and fungi

A clear differentiation in substrate utilisation pattern was visible between

treatments (Fig. 9.1A and 9.1B). No difference was observed in bacterial or

fungal substrate utilisation patterns (BiologTM ecoplate) between treated field

soils and the virgin brigalow soil until 72 h of incubation (Fig. 9.1). Although the

metabolic activity of fungi in the control treatment seems slightly higher than the

other treatments, it was not significantly different. The rate of substrate utilisation

was higher in brigalow soil than in cultivated soil, suggesting higher metabolic

activities of both bacterial and fungal communities in the brigalow soil compared

to those of other treatments.

Significant differences among treatments in average utilisation of

substrate guilds were evident within BiologTM (Ecoplate and FF microplate) data

(Fig. 9.2). Samples from the brigalow treatment averaged higher utilisation of

carbohydrates in the Ecoplate than those of other treatments while biology boom

spray samples averaged higher utilisation of miscellaneous substrates than the

feather-top Rhodes treatment. Similarly, brigalow samples averaged higher

utilisation of carbohydrates than biology direct injection, and higher utilisation of

polymers compared to other treatments, except feather-top Rhodes, in the FF

microplate (Fig. 9.2B).

229
Chapter 9: Compost extract and soil microbial activity

0.30 A
Control
Best-bet conventional
Biology boom spray
0.25
Biology direct injection
Best-bet biology
Green manure
Feather-top Rhodes
0.20
Brigalow soil

AWCD 570 nm
0.15

0.10

0.05

0.00
0 24 48 72 96 120 144 168

0.30
Control B
Best-bet conventional
Biology boom spray
0.25 Biology direct injection
Best-bet biology
Green manure
Feather-top Rhodes
0.20 Brigalow soil
AWCD 490 nm

0.15

0.10

0.05

0.00
0 24 48 72 96 120 144 168
Time (h)

Figure 9. 1. Kinetics of average well colour development (AWCD) of (A) bacterial (570 nm) and
(B) fungal (490 nm) communities originating from soil samples collected from the trial site during
the dry-winter season at Baralaba. Values are mean ± SE (n = 3) for all treatments except for
green manure, feather-top Rhodes and brigalow soil (n = 2).

230
Chapter 9: Compost extract and soil microbial activity

Figure 9. 2. Average carbon substrate utilisation for different substrate categories in terms of (A)
32 (Biolog Ecoplate) and (B) 96 (Biolog FF microplate) different food sources. Samples were
collected from 72 h incubations of both Biolog plate types and utilisation was measured as the
average optical density across all substrates within each guild (i.e. across 7 different
carbohydrates, 9 carboxylic acids, 4 polymers, 6 amino acids, 2 amines/amides and 3
miscellaneous for Ecoplate, and 44 different carbohydrates, 17 carboxylic acids, 5 polymers, 13
amino acids, 6 amines/amides and 10 miscellaneous) for FF microplate.

Principal components analysis (PCA) of the Biolog Ecoplate assay data

explained 43% and 20% of the total variance in the first two principal

components (Fig. 9.3A). A plot of PC1 against PC2 revealed that the bacterial

communities present in the brigalow soil were quite different to that of the

cropping soils (Fig. 9.3A). The communities in the cropping treatments were

231
Chapter 9: Compost extract and soil microbial activity

clustered, although a decrease in principal component 2 characterised some of the

best-bet conventional and feather-top Rhodes treatment replicates.

The PCA of FF microplate data explained 38% (PC1) and 21% (PC2) of

the total variance (Fig. 9.3B). In a PC2 against PC1 plot, the brigalow soil

samples were separated from the main cluster of cropped soil samples, except for

one sample from the control group (with an extreme PC2 value) and one sample

from the best-bet biology treatment (extreme PC1 value) (Fig. 9.3B).

Figure 9. 3. Principal components analysis of optical densities produced by microbial activities of


soil samples collected from the trial site at Baralaba consuming (A) 32 (Biolog Ecoplate) and (B)
96 (Biolog FF microplate) different food sources; representing bacterial and fungal diversity,
respectively. Values are mean (n = 3) for all treatments except for green manure, feather-top
Rhodes and brigalow soil (n = 2).

232
Chapter 9: Compost extract and soil microbial activity

3.3. PCR-DGGE analysis for bacteria and fungi

Extracted soil DNA from each treatment (n=2) were run on the DGGE

gels to obtain microbial community patterns. Replicate variability was higher in

fungal compared to bacterial communities.

A UPGMA (Unweighted Pair Group Method with Arithmetic mean

algorithm) dendrogram was based on the DGGE generated DNA profiles (Fig.

9.4). A shift in microbial communities occurred between seasons (Fig. 9.4),

although the change was more marked in the bacterial than the fungal community

(Fig. 9.4).

The UPGMA dendrogram for bacterial community formed four distinct

clusters, where best-bet biology treatment grouped with control in either season

whereas brigalow grouped with biology direct injection, biology boom spray and

best-bet conventional treatments (Fig. 9.4A). Similarly, feather-top Rhodes and

green manure treatments were clustered together showing relatedness of bacterial

communities between the two treatments. Unlike the PCA analysis of the Biolog

(Ecoplate) substrate utilisation data, the UPGMA dendrogram did not separate

the brigalow communities with those of the field treated soils.

Likewise, the dendrogram for fungal communities formed five clusters

where best-bet biology grouped with control samples collected in the wet-

summer season whereas in the dry-winter season, fungal communities of biology

boom spray, biology direct injection, best-bet biology and control treatments

clustered together (Fig. 9.4B). Brigalow soil formed a separate cluster whereas

feather-top Rhodes and green manure clustered together. The best-bet

conventional treatment, however, showed replicate variability and formed a

separate cluster.

233
Chapter 9: Compost extract and soil microbial activity

Figure 9. 4. Unweighted Pair Group Method with Arithmetic mean algorithm dendrogram
constructed from the DGGE profiles of (A) bacterial 16S rDNA and (B) fungal ITS genes
amplified from duplicate DNA samples of soils collected in the dry-winter season, except control
(I) and best-bet biology (I) which were additionally collected in the previous wet-summer season
(highlighted in the figure), from the trial site and the nearby virgin brigalow soil at Baralaba. The
scale bar represents percent similarity.

The UPGMA dendrogram generated for bacterial communities also

showed that best-bet biology and biology direct injection treatments were

clustered with approximately 60% homology forming a sub group (Fig. 9.4A).

234
Chapter 9: Compost extract and soil microbial activity

The fungal dendrogram showed approximately 65% homology between best-bet

biology and biology direct injection treatments (Fig. 9.4B). Both of these

treatments involved compost extract applied via liquid injection during planting,

but with and without commercial microbial inoculant (Twin N ‗diazotrophs‘),

respectively. The results of the PCR-DGGE analysis (Fig. 9.4A) were consistent

with the principal components analysis (BiologTM) data, in that the bacterial

community structure of the biology direct injection and best-bet conventional

treated soils demonstrated relatedness (Fig. 9.3A). The fungal community

structure of soils treated with biology boom spray, biology direct injection and

best-bet biology treatments showed relatedness to one another (Fig. 9.3B and

9.4B), but not to the degree shown by the bacterial communities.

Bacterial community diversity of best-bet biology as denoted by Shannon-

Weaver diversity index (H‘) was signifcantly higher compared to the best-bet

conventional treatment (Table 9.3). Although the UPGMA dendrogram for either

bacteria and fungi showed considerable differences, no changes in their diversity,

evenness and richness were evident in the control treatments when compared

between the two seasons (Table 9.3). Nevertheless, significantly higher bacterial

diversity was found in the best-bet biology system in the dry-winter compared to

the wet-summer season (Table 9.3).

235
Chapter 9: Compost extract and soil microbial activity

Table 9. 3. Shannon-Weaver Diversity Index (H‘), Equitability Index (J) and number of bands (S)
for bacteria and fungi present in soil samples collected in the dry-winter season from plots treated
with different management practices including compost extract treatments with two methods of
application at Baralaba trial site. All soil samples were collected in the dry-winter season, except
control (I) and best-bet biology (I) which were also collected in the previous wet-summer season.
Within columns, mean ± SE (n = 2) with the same letter are not significantly different according
to Tukey‘s test (p< 0.05).
Bacteria Fungi
Treatments H' J S H' J S
Control 2.26 ± 0.10b 0.88 ± 0.04a 13 ± 0.00b 2.88 ± 0.04cd 0.81 ± 0.01bc 35 ± 0.50cd
Best-bet biology 2.74 ± 0.10c 0.91 ± 0.01a 21 ± 1.50c 2.46 ± 0.02abc 0.74 ± 0.01abc 28 ± 0.50ab
Best-bet conventional 2.33 ± 0.01b 0.92 ± 0.01a 13 ± 0.50b 2.65± 0.21abcd 0.77 ± 0.04abc 31 ± 3.00abc
Biology boom spray 2.39 ± 0.03bc 0.92 ± 0.00a 14 ± 0.50b 2.32 ± 0.11a 0.71 ± 0.03a 26 ± 0.00ab
Biology direct injection 2.39 ± 0.01bc 0.96 ± 0.01a 12 ± 0.00b 2.54 ± 0.05abcd 0.75 ± 0.01abc 29 ± 1.00abc
Green manure 2.43 ± 0.08bc 0.92 ± 0.01a 14 ± 1.00b 2.79 ± 0.03bcd 0.80 ± 0.01abc 32 ± 0.00bcd
Feather-top Rhodes 2.32 ± 0.13b 0.90 ± 0.02a 13 ± 1.00b 2.37 ± 0.07ab 0.73 ± 0.01ab 26 ± 1.50a
Brigalow soil 2.35 ± 0.01bc 0.94 ± 0.01a 12 ± 0.00b 2.96 ± 0.02d 0.82 ± 0.00bc 38 ± 0.50d
Control (I) 2.35 ± 0.08bc 0.93 ± 0.01a 13 ± 0.50b 2.94 ± 0.01d 0.83 ± 0.00c 35 ± 0.00cd
Best-bet biology (I) 1.86 ± 0.02a 0.95 ± 0.01a 7 ± 0.00a 2.82 ± 0.02cd 0.80 ± 0.00abc 34 ± 0.00cd

No differences were recorded among treatments in terms of evenness (J)

of bacterial communities, however richness (S) in bacterial communities was

higher in best-bet biology than that of other treatments including biology direct

injection. Both biology direct injection and best-bet biology received similar

treatments but best-bet biology was treated with Twin N inoculant (free-living

rhizosphere N2 fixers) prior to the onset of flowering (Table 9.3). Bacterial

community diversity (H‘) and richness (S) in best-bet biology treatment were

higher in the wet-summer compared to the dry-winter. The lack of difference in

evenness scores among all treatments indicates the uniformity of band intensities

in the DGGE profiles for bacteria (Table 9.2). A shift in soil bacterial

communities was produced by best-bet biology treatment, in contrast to the best-

bet conventional treatment, although fungal communities of those treatments

remain unchanged.

The ITS-DGGE data showed no alterations in the fungal diversity and

evenness of control and biological treatments between the two seasons, however

fungal richness was higher in best-bet biology treatment in the wet-summer than

236
Chapter 9: Compost extract and soil microbial activity

in the dry-winter season (Table 9.3). ITS-DGGE profiles revealed the presence of

few common bands in all soil treatments along with the virgin brigalow land with

considerable variation in relative intensity of common bands between treatments.

In the dry-winter season, best-bet biology, biology boom spray and feather-top

Rhodes treatments demonstrated lower fungal diversity as assessed by Shannon-

Weaver diversity values compared to the virgin brigalow soil, and even compared

to the control (Table 9.3).

The fungal communities in the soil samples, however, showed greater

diversity (H‘) and relative abundance or richness (S) of fungi, as indicated by the

higher number of bands, than equivalent bacterial communities irrespective of the

treatment effects (Table 9.3).

4. Discussion

4.1. Seasonal variation

The summer season enjoys higher temperatures and rainfall, and thus soil

moisture, and higher plant biomass, conditions expected to favour soil microbial

activity and mineralisation. For example, Kaiser et al. (1995) also observed

temporal variation in microbial biomass C of -15% to +15% of the annual mean

in winter and summer seasons, respectively, for an arable soil (luvisol). In the

current experiment (PWP), the soil field capacity of the vertisol soil was 33%

w/v, while permanent wilting point is associated with a moisture content of 19%

w/v (Chinn & Pillai, 2008). In the current study, the moisture content of the

surface (0-10 cm) soil in the wet season was around 13% w/v, or well below

PWP. The moisture content of the soil collected in the dry season was adjusted to

a similar moisture level for a 5 d incubation period. However, plate counts of

237
Chapter 9: Compost extract and soil microbial activity

bacteria, fungi, and microbial biomass C (in average) were higher in the wet-

summer than the following dry-winter (Table 9.2).

Although higher plate counts and microbial biomass were found in soils

collected in summer season, it was not reflected in the microbial activity. Soil

microbial activity and soil respiration (both basal and substrate induced) were

higher in soil collected from winter compared to that from the previous summer

season (Table 9.2). It could be that the fungal spores were present in resting

stages in soil during summer given that the moisture content of the soil collected

was well below PWP. Although winter soil was wet up to this level, wetting was

never uniform. Therefore, it is possible that there were pockets of soil at water

potentials closer to zero, supporting microbial activity.

4.2. Comparison of brigalow and cultivated soils

The brigalow forest soil was significantly higher in nitrate, moisture

content, bacterial and fungal populations, microbial biomass C, substrate induced

respiration and microbial activity in the dry-winter season compared to the

cultivated soil (Table 9.2). This could be ascribed to the fact that the forest soil

has benefit of continual vegetation and mulch cover, with fire acting in removal

of organic matter and deposition of carbon. There are many studies documenting

changes in soil over time since brigalow clearing (Radford et al., 2007; Solomon

et al., 2000; Thornton et al., 2007). Collard and Zammit (2006) reported reduced

levels of total organic as well as labile carbon in intensively used agricultural

land compared to the remnant brigalow soil. Labile carbon and microbial biomass

in soil are noted to be strongly correlated, with microbial biomass linked to the

rate of turnover of organic matter (Bell et al., 1998). Spaccini et al. (2001) also

238
Chapter 9: Compost extract and soil microbial activity

demonstrated that land cultivation reduced soil aggregate stability, with forest

soil richer in organic carbon.

The PCA of the Biolog data of the present study (Fig. 9.3) suggests that

original bacterial communities of the virgin brigalow soil have been shifted by

cultivation practices. However, culture based methods, such as Biolog neglect the

presence of non-culturable microorganisms, and as such are not totally true

guides to the microbial species composition of the sample (Widmer et al., 2001).

The culture limitation of the Biolog technique is likely to affect slow growing

organisms, and organisms inhabiting different ecological niches to that tested in

the Biolog procedure (e.g. obligate anaerobes). Smalla et al. (1998) characterised

the microbial communities of activated sludge samples using Biolog GN plates

followed by DGGE and TGGE techniques in order to explore which microbial

populations of the communities generate the Biolog patterns. Using the molecular

techniques, they revealed that the highly adaptive and fast-growing communities

of bacteria, dominant in the Biolog wells, were responsible for the pattern

observed. Therefore, the PCR-DGGE technique was applied in the current study

in order to verify the results obtained from Biolog.

DGGE profiles for the treated plots revealed distinct shifts between

biological and conventional treatments, however few bands were present in all

treatments, and some bands were unique either to the cultivated land or to the

virgin brigalow soil. No difference between the bacterial communities in

brigalow and the field-treated soils as suggested by the UPGMA dendrogram and

the DGGE profile indicate that both cultivated and brigalow soils share the

common bacterial community structure. Although not performed in this study, the

DGGE generated bands can be excised from the gels, sequenced, and analysed

239
Chapter 9: Compost extract and soil microbial activity

phylogenetically to species level through comparison with sequences available in

GeneBank database (e.g. www.ncbi.nlm.nih.gov) (Jing et al., 2010).

In the dry-winter season, fungal diversity as assessed by Shannon-Weaver

diversity was higher in the virgin brigalow soil compared to best-bet biology,

biology boom spray and feather-top Rhodes treatments (Table 9.3). Higher fungal

diversity in the brigalow soil may be related to the presence of the recalcitrant

forest litter, compared to the plant debris of cultivated soils (Lauber et al., 2008).

This is consistent with the results of Biolog where brigalow sample performed

better on polymers in FF microplates (Fig. 9.2).

The greater diversity of soil fungal communities than equivalent bacterial

communities in brigalow soil, noted in this study (Table 9.3), is in accordance

with the findings of Bailey et al. (2002), who also observed greater fungal to

bacterial biomass, as measured by phospholipid fatty acid analysis (PLFA), in

forest soil compared to the agriculturally manipulated soils. Unfavorable climatic

conditions, such as insufficient rainfall might have stimulated production of

fungal spores, which can remain viable in dry soil for longer periods (Elliot et al.,

2002). In the biology boom spray treatment, the compost extract is sprayed on the

surface of the soil which could be one of the reasons why the soil exhibited the

least fungal diversity, while the bacterial communities remained uniform to the

other treatments (Table 9.3).

4.3. Effects of management practices on soil properties

Both Twin-N inoculant (rhizosphere N2 fixers) and green manuring crops

could have contributed to a significant increase in ammonium concentration in

biology direct injection and green manure treatments (Table 9.2) during summer.

240
Chapter 9: Compost extract and soil microbial activity

Similarly, the granular fertiliser ‗Starter Z‘ (19.5% P) could be the source of

higher phosphate concentration in best-bet conventional treated soils compared to

other treatments in the same season.

Increased soil respiration is a direct indicator of increased microbial

activity (Gomez et al., 2001; Parkin et al., 1996). In the present study,

significantly higher basal respiration was noted in best-bet conventional and

green manure, compared to feather-top Rhodes, whereas substrate induced

respiration was significantly higher in biology direct injection and feather-top

Rhodes compared to the control in the summer season (Table 9.2). This could be

due to stimulation of soil microbiota and increase in microbial biomass with the

addition of green manures (Stark et al., 2007) and mineral fertilisers (Chu et al.,

2007), or it could simply be due to the synergistic effect of soil indigenous and

exogenous microbiota supplied through the organic amendments and

biofertilisers (Prasanna et al., 2008).

5. Conclusions

The observation that soil microbial diversity is different in undisturbed

brigalow soil to cultivated soils is interesting, but not unexpected. After all, the

above-ground community structure is also quite different to that of cultivated

land.

The agricultural management practices applied in this study demonstrated

little impact on soil microbial diversity relative to a seasonal effect. However,

compost extract application in soil along with foliar applications of microbial

inoculant and nutrients improved bacterial communities in soil compared to the

conventional treatment.

241
Chapter 9: Compost extract and soil microbial activity

Future exploration of the practical benefits of compost extracts should

focus on the status of key microbes – key to various soil processes, for example,

Rhizobium for legume nodulation and N2 fixation, rhizobacteria (such as Bacillus,

and Nitrobacter) for organic phosphate mineralisation and inorganic phosphate

solubilisation, making nitrogen and phosphorus available to plants. Other groups

such as ammonia oxidizers or nitrifiers, sulphur oxidizing bacteria, mycorrhizal

fungi, which are considered important in nutrient cycling, nutrient transport and

chelating of toxic elements should also be considered for future work. Some

agronomic information to suggest some target species which help supply these

nutrients limiting for crop production in central Queensland, where ferrosol is

generally limited in P, should also be explored.

242
Chapter 10: Summary and future directions

Chapter 10

Summary and future directions

Microbially enhanced compost extract (compost tea), obtained after

incubating compost in water with microbial food sources, is claimed to have

benefits in improving soil and crop ‗health‘ (Ingham, 2005), but there is relatively

little in the scientific literature on the effectiveness and mode of action of this

product (as reviewed in Chapter 1). The following hypotheses were established as

to the benefit of compost extract to plant growth: (a) direct improvement of plant

nutrition; (b) improvement of plant nutrition through an indirect action, via

microbial activities in mineralising or solubilising of organic and inorganic

components of the soil; (c) acceleration of litter decomposition; (d) bio-protection

against fungal pathogens (e.g. Fusarium oxysporum f. sp. lycopersici); and (e)

protection of root growth against high levels of toxic soil substances (e.g. Al+++),

via chelation of Al ions by humic acids or uptake by microbes.

The results of this thesis indicated that compost extracts (a) improved

plant growth via direct nutrient addition, (b) but not via mineralisation or

solubilisation of organic or inorganic matter in soil, and (c) accelerated litter

decomposition to some extent, (d) reduced root fungal diseases, and (e) reduced

243
Chapter 10: Summary and future directions

aluminium toxicity possibly via the action of auxin or gibberellic acid-like

growth regulators.

The effect of compost extract on plant growth in the tomato and sorghum

pot experiments was largely attributed to the nutrients inherent in the high dose of

compost extract applied (equivalent to 34,000 L ha-1). Of course, field application

of such large quantities is not practical in most production systems. However,

such compost extracts may find as an organic liquid fertiliser, for example,

‗organic hydroponics‘, combining two higher values, but normally diametrically

opposed categories, in the supply chain.

There was no evidence of an increase in solubilisation of soil minerals or

mineralisation of soil organic matter, sufficient to impact on plant nutrition.

Compost extracts sprayed onto leaf litter did result in a long term respiration rate

increase, however the impact on litter breakdown was limited over the 168 day

trial. Acceleration of stubble breakdown is a logical application of compost

extract. Use of biological sprays to assist stubble breakdown is well documented

(e.g. Freeman et al., 1948; Greathouse, 1949; Nilsson, 1974; Nilsson & Ginns,

1979), and several commercial providers exist (e.g. supplying cellulose digesting

fungi, specifically Chaetomium brasiliense and Trichoderma harzianum;

www.bionutrient.com.au/products/full-product-list-46/stubble-digest-57.htm).

The relatively poor rate of sugarcane litter breakdown observed in the present

study might be addressed by using compost based on sugarcane trash, or by

species specific enrichment of the compost extract. Attention could also be given

to the degradation environment (e.g. moisture levels; trash particle size; trash

turning). This application is worthy of further attention, being of significance to

244
Chapter 10: Summary and future directions

the Australian sugarcane industry as it adopts the practice of green cane

harvesting.

Further study using lower dilutions of compost extract with respect to the

amelioration of aluminium toxicity is warranted, and on the accumulation of

humates in the soil from successive additions of compost extract. The observed

effect of amelioration of toxicity (200 μM) on maize roots following imbibitions

in compost extract hints at the involvement of auxin or gibberellic acid-like

growth regulators. Chemical identification of these agents would be useful, in

order to allow compost extracts to be tailored to favour this function.

Compost extracts have been reported to be of use in suppressing plant

diseases. Although the results of the present studies conducted in vitro on the

efficacy of compost extracts in suppressing fungal wilt diseases were promising,

the effect of compost extracts on disease suppression in tomato crops in

hydroponics culture was not consistent for the three tested pathogens. Future

effort should focus on identifying techniques to reduce such inherent variability,

which may require the identification of suppressive microbiota, ensuring their

survival during the production process, and ensuring the delivery of desired

microbes onto the roots or foliage of plants (e.g. through various application

techniques along with wetting agents, spray adjutants, and emulsifiers).

Integrated approaches such as multiple cropping, crop rotations and fallow

cropping should be used along with the use of compost extracts at the field scale,

the latter needing further optimisation to control such diseases effectively. It

should, however, be understood that compost extract is not a panacea for all

diseases, and therefore growers need to employ integrated plant health

management tools in parallel to the use of compost extracts.

245
Chapter 10: Summary and future directions

Rumen based compost extracts were compared with two commercial

products with respect to nutritional and microbial characteristics. The question of

whether a relatively uncontrolled microbial species mix (as present in enhanced

compost extract) or controlled single or multiple species tailored to specific

applications should be recommended for agronomic use remains to be answered.

Certainly, the Rhizobium industry has developed around the use of single strains

targeted to specific crops (in Australia) (Brandon et al., 1998) or a mix of a

number of strains of the one species, for use across a number of crop types (in the

USA) (Stacey & Upchurch, 1984). Similarly, relatively pure cultures of

microorganisms, such as Azotobacter, Trichoderma, and Azospirillum are

commercially available as ‗biofertilisers‘ (e.g.

www.biofertilizer.com/organic_garden_biofertilizers/biofertilizer). Arguably, an

extract could be tailored for an application if it were inoculated with specific

microbial species selected for a specific task (litter breakdown, disease

protection, phosphate solubilisation, or similar). Indeed, this is the approach taken

by SmartBugs Pty Ltd, who markets a freeze-dried inoculum of beneficial

microbes (e.g. Azotobacter spp., Azospirillum spp., Bacillus spp., Trichoderma

spp., Cellulomonas spp.) instead of compost, for use in production of ‗compost

tea‘. Conversely, the ‗low tech‘ option of producing a compost ‗tea‘ from locally

produced compost holds appeal to those who favour a ‗natural‘ approach.

The compost extract incubation process attempts to idealise microbial

growth conditions - but there is the risk of growth of pathogenic microorganisms.

Although not investigated in this study, human-health-related issues of compost

extracts is also of concern and should also be addressed in future studies.

246
Chapter 10: Summary and future directions

In this thesis, an open air incubation system was characterised in terms of

the incubation history of pH, EC, temperature and bacterial-fungal population

level. This type of basic description could be undertaken to compare the range of

commercially available incubators that have become available (e.g. fitted with

advanced aerators). Further study on assessing and increasing the storability or

shelf-life of compost extracts for storage to allow for long distance transport is

also recommended.

The quality of the compost is one of the determinants of the quality of the

compost extract. Vermicomposting, followed by separation and collection of

leachate from the solid waste, was seen in the present study to increase the

microbial biomass of the rumen compost. Further, liquid incubation of composts

to produce an extract altered their nutrient and microbial composition with

highest functional diversity observed in vermicast leached extract compared to

the original composts. Thus, there is good scope for the utilisation of earthworms

as a means of improving the quality of the inoculum.

In the field study, although little effect was found on the fungal

communities, total and active bacterial communities were enhanced by compost

extract application compared to the other treatments. It is likely that compost

extracts will find an expanded role in broadscale agriculture.

The ability of these liquid extracts to favourably shift the activity and

diversity of desirable microorganisms in field soil, when applied at low

application rates (e.g. 200 L ha-1) is another interesting area for consideration.

Application of key ‗missing‘ microbiota can have a large scale impact from a

small scale application (e.g. as with rhizobia or mycorrhizae), but addition of

microbiota already present in the soil is unlikely to have such an impact. Further

247
Chapter 10: Summary and future directions

research on field survival and establishment of such microbiota added to the soil

in compost extracts is warranted, combined with consideration of application

frequency, rate, and method (liquid injection vs. fertigation or foliar spray, seed

priming vs. soil drench). Attention should be given to studies on combinations of

potential plant growth promoting microorganisms for improved results. Perhaps

ultimately there could be a small set of ‗indicator‘ species one would quantify as

an index of a given soil‘s capacity for mineralisation, phosphate solubilisation,

and plant disease protection.

Other potential areas for further study include:

1. the ability of compost extract to alter microbial populations on the

phylloplane of plants following low rates (e.g. 20-50 L ha-1) of foliar

application.

2. the ability of compost extract to improve plant nutrition through

mineralisation of the soil organic fraction, especially to P nutrition, through

P-solubilisation by the action of bacteria and mycorrhizae.

3. the ability of compost extract to result in increased N2 fixation in the soil by

free living soil bacteria.

4. the role of compost extract in detoxifying organic compounds, for example,

pesticides, and inorganic ions other than Al+++ (e.g. Mn) in acid soils by

organic acids or sequestration by microbiota.

5. the effect of compost extract in improving soil structure through direct

addition of humates to the soil and through addition of organic matter via

partial decomposition of plant residues.

248
Chapter 10: Summary and future directions

6. the role of compost extract in improving water holding capacity of soil

through improvement in soil structure and through addition of organic matter

via partial decomposition of plant residues.

7. the role of compost extract with respect to increasing plant resistance to

insects and insect deterrence or control.

The findings of this research have provided an insight to the

users/producers of microbially enhanced compost extract or ‗compost tea‘. At the

same time, it has generated more questions to be answered and will hopefully

encourage future researchers to investigate the other possible roles of microbially

enhanced compost extracts.

249
References

References

Abbott, L.K., Murphy, D.V. 2003. Soil Biological Fertility: A Key to Sustainable

Land Use in Agriculture, Dordrecht; Boston: Kluwer Academic

Publishers, pp. 264.

Adam, G., Duncan, H. 2001. Development of a sensitive and rapid method for the

measurement of total microbial activity using fluorescein diacetate (FDA)

in a range of soils. Soil Biol. Biochem., 33, 943-951.

Adams, J.D.W., Frostick, L.E. 2009. Analysis of bacterial activity, biomass and

diversity during windrow composting. Waste Management, 29, 598-605.

Adams, P.B. 1990. The potential of mycoparasites for biological control of plant

diseases. Annu. Rev. Phytopathol., 28, 59-77.

Adegunloye, D.V., Adetuyi, F.C., Akinyosoye, F.A., Doyeni, M.O. 2007.

Microbial analysis of compost using cowdung as booster. Pakistan J.

Nutr., 6, 506–510.

Aibuzio, A., Ferrari, G., Nardi, S. 1986. Effects of humic substances on nitrate

uptake and assimilation in barley seedlings. Can. J. Soil Sci., 66, 731-736.

Aira, M., Dominguez, J. 2008. Optimising vermicomposting of animal wastes:

Effects of rate of manure application on carbon loss and microbial

stabilisation. J. Environ. Management, 88, 1525-1529.

Al-Mughrabi, K.I., Berthelerne, C., Livingston, T., Burgoyne, A., Poirier, R.,

Vikram, A. 2008. Aerobic compost tea, compost and a combination of

both reduce the severity of common scab (Streptomyces scabies) on

potato tubers. J. Plant Sci., 3(2), 168-175.

250
References

Albanell, E., Plaixats, J., Cabrero, T. 1988. Chemical changes during

vermicomposting (Eisenia fetida) of sheep manure mixed with cotton

industrial wastes. Biol. Fertil. Soils, 6, 266-269.

Alexander, M. 1994. Biodegradation and Bioremediation. Academic Press, New

York.

Allard, A.S., Neilson, A.H. 1997. Bioremediation of organic waste sites: a critical

review of microbiological aspects. Int. Biodeterioration &

Biodegradation, 39(4).

Allen, D.J., Ort, D.R. 2001. Impact of chilling temperatures on photosynthesis in

warm-climate plants. Trends Plant Sci., 6, 36–42.

Alms, M. 2002. Castings tea: Panacea or fantasy? in: Vermico’s Bimonthly

Newsletter, Vol. 7.

Altomare, C., Norvell, W.A., Björkman, T., Harman, G. 1999. Solubilization of

phosphates and micronutrients by the plant growth-promoting and

biocontrol fungus Trichoderma harzianum Rifai 1295-22. Appl. Environ.

Microbiol., 65, 2926–2933.

An, D.S., Im, W.T., Yang, H.C., Kang, M.S., Kim, K.K., Jin, L., Kim, M.K., Lee,

S.T. 2005. Cellulomonas terrae sp. nov., a cellulolytic and xylanolytic

bacterium isolated from soil. Int. J. Syst. Evol. Microbiol., 55(4), 1705-

1709.

Anderson, J.P.E. 1982. Soil respiration. Second ed. in: Methods of Soil Analysis,

Am. Soc. Agron. Madison, WI, pp. 831–871.

Anderson, T.-H. 2003. Microbial eco-physiological indicators to assess soil

quality. Agric. Ecosyst. Environ., 98, 285-293.

251
References

Andrew, J.H. 1993. Compost extracts and the biological control of foliar plant

disease.

Antonio, G.F., Carlos, C.R., Reiner, R.R., Miguel, A.A., Angela, O.L.M., Cruz,

M.J.G., Dendooven, L. 2008. Formulation of a liquid fertiliser for

sorghum (Sorghum bicolour (L.) Moench) using vermicompost leachate.

Bioresour. Technol., 99, 6174-6180.

Arthur, G.D., A.K., J., J., V.S. 2001. The release of cytokinin-like compounds

from Gingko biloba leaf material during composting. Environ. Exp. Bot.,

45, 55-61.

AS4454. 1999. Australian standard - composts, soil conditioners and mulches

standards, Association of Australia Homebush, NSW.

Atagana, H.I. 2009. Biodegradation of PAHs by fungi in contaminated-soil

containing cadmium and nickel ions. Afr. J. Biotechnol., 8(21), 5780-

5789.

Avnimelech, Y., Eilat, R., Porat, Y., Kottas, P.A. 2004. Factors affecting the rate

of windrow composting in field studies. Compost sci. util., 12(2), 114-

118.

Bailey, K.L., Lazarovits, G. 2003. Suppressing soil-borne diseases with residue

management and organic amendments. Soil Tillage Res., 72, 169-180.

Bailey, V.L., Smith, J.L., Bolton, H. 2002. Fungal-to-bacterial ratios in soils

investigated for enhanced C sequestration. Soil Biol. Biochem., 997–1007,

34.

Ball-Coelho, B., Tiessen, H., Stewart, J.W.B., Salcedo, I.H., Sampaio, E.V.S.B.

1993. Residue management effects on sugarcane yield and soil properties

in northeastern Brazil. Agron. J., 85, 1004-1008.

252
References

Ball-Coelho, B.R., Sampaio, E.V.S.B., Tiessen, H., Stewart, J.W.B. 1992. Root

dynamics in plant and ratoon crops of sugarcane. Plant Soil, 142, 297-

305.

Barzengar, A.R., Yousefi, A., Daryashenas, A. 2002. The effect of addition of

different amounts and types of organic materials on soil physical

properties and yield of wheat. Plant Soil, 247, 295-301.

Begon, M., Harper, J.L., Townsend., C.R. 1990. Ecology: individuals,

populations, and communities. Blackwell Sci. Pub., 945.

Bell, M.J., Moody, P.W., Connolly, R.D., Bridge, B.J. 1998. The role of active

fractions of soil organic matter in physical and chemical fertility of

Ferrosols. Aust. J. Soil Res., 36, 809-819.

Bell, M.J., Stirling, G.R., Pankhurst, C.E. 2007. Management impacts on health

of soils supporting Australian grain and sugarcane industries. Soil Tillage

Res., 97, 256-271.

Bennet, R.J., Breen, C.M. 1991. The aluminium signal: new dimensions of

aluminium tolerance. Plant Soil, 134, 153-166.

Bernal-Vicente, A., Ros, M., Tittarelli, F., Intrigliolo, F., Pascual, J.A. 2008.

Citrus compost and its water extract for cultivation of melon plants in

greenhouse nurseries. Evaluation of nutriactive and biocontrol effects.

Bioresour. Technol., 99, 8722–8728.

Berner, A., Hildermann, I., Fließbach, A., Pfiffner, L., Niggli, U., Mader, P.

2008. Crop yield and soil fertility response to reduced tillage under

organic management. Soil Tillage Res., 101, 89–96.

Bess, V.H. 2000. Understanding compost tea. BioCycle, 41 (10), 71-72.

253
References

Bingrui, J., Guangsheng, Z. 2009. Integrated diurnal soil respiration model during

growing season of a typical temperate steppe: Effects of temperature, soil

water content and biomass production. Soil Biol. Biochem., 41, 681–686.

Blair, N., Crocker, G.J. 2000. Crop rotation effects on soil carbon and physical

fertility of two Australian soils. Aust. J. Soil Res., 38, 71–84.

Bloem, J., Bolhuis, P.R., Veninga, M.R., Wieringa, J. 1995. Microscopic methods

for counting bacteria and fungi in soil. in: Methods in Applied Soil

Microbiology and Biochemistry, (Eds.) K. Alef, P. Nannipieri, Academic

Press. London, pp. 162–191.

Bloem, J., Lebbink, G., Zwart, K.B., Bouwman, L.A., Burgers, S.L.G.E., de Vos,

J.A., de Ruiter, P.C. 1994. Dynamics of microorganisms, microbivores

and nitrogen mineralisation in winter wheat fields under conventional and

integrated management. Agric. Ecosys. Environ., 51(1-2), 129-143.

Bourn, D., Prescott, J. 2002. A comparison of the nutritional value, sensory

qualities, and food safety of organically and conventionally produced

foods. Crit. Rev. Food Sci. Nutr., 42(1), 1–34.

Bourne, M., Nicotra, A.B., Colloff, M.J., Cunningham, S.A. 2008. Effect of soil

biota on growth and allocation by Eucalyptus microcarpa. Plant Soil,

305(1-2), 145-156.

Brandon, N.J., Date, R.A., Clem, R.L., Robertson, B.A., Graham, T.W.G. 1998.

Growth responses of Desmanthus virgatus to inoculation with Rhizobium

strain CB3126. II. A field trial at 4 sites in south-east Queensland.

Tropical Grasslands, 32, 20-27.

Brinton, W.F. 1995. The control of plant pathogenic fungi by use of compost

teas. Biodynamics, 12-15.

254
References

Brinton, W.F., Storms, P., Evans, E., Hills, J. 2004. Compost teas: Microbial

hygiene and quality in relation to method of preparation. J. Biodynamics,

1-9.

Brinton, W.F., Tranker, A., Droffner, M. 1996. Investigation into liquid compost

extracts. BioCycle, 37(11), 68-70.

Bryant, T. 2008. When under attack, plants can signal microbial friends for help.

in: Science News, University of Delaware.

Buckley, D.H., Schmidt, T.M. 2001. The structure of microbial communities in

soil and the lasting impact of cultivation. Microbial Ecol., 42, 11–21.

Bulluck, L.R., Brosius, M., Evanylo, G.K., Ristaino, J.B. 2002. Organic and

synthetic fertility amendments influence soil microbial, physical and

chemical properties on organic and conventional farms. Appl. Soil Ecol.,

19, 147–160.

Buswell, J.A., Eriksson, K.-E.L. 1994. Effect of olignin-related phenols and their

methylated derivatives on the growth of eight white-rot fungi. World J.

Microbiol. Biotech., 10, 169-174.

Campbell, A. 2006. Overview of compost tea use in New South Wales. Recycled

Organics Unit, Department of Environment and Conservation, The

University of New South Wales.

Canullo, G.H., Rodrıguez-Kabana, R., Kloepper, J.W. 1992. Changes in the

populations of microorganisms associated with the application of soil

amendments to control Sclerotium rolfsii. Plant Soil, 144, 59–66.

Carpenter, L. 2005. Diving into compost tea. BioCycle, 46(7), 61.

Casagrande, A.A., Melo, W.J.D., Pereira, R., Mutton, M.A., Barbosa, J.C.,

Campos, M.S. 1995. Influence of residues from mechanical harvesting

255
References

and vinasse on soil nitrogen in sugarcane ratoon crops. Proc. Int. Soc.

Sugarcane Technol., 22, 67-69.

Casenave de Sanfilippo, E., Arguello, J.A., Abdala, G., Orioli, G.A. 1990.

Content of auxin-, inhibitor- and gibberellin-like substances in humic

acids. Biologia Plantarum, 32, 346–351.

Castella, G., Bragulat, M.R., Rubiales, M.V., Caba˜nes, F.J. 1997. Malachite

green agar, a new selective medium for Fusarium spp. Mycopathol., 137,

173–178.

Chef, D.G., Hoitink, H.A.J., Madden, L.V. 1983. Effects of organic components

in container media on suppression of Fusarium wilt of crysanthemum and

flax. Phytopathol., 73, 279-281.

Chinn, C., Pillai, U.P.P. 2008. Self-repair of compacted vertisols from Central

Queensland, Australia. Geoderma, 144(3-4), 491-501.

Chu, H., Lin, X., Fujii, T., Morimoto, S., Yagi, K., Hu, J., Zhang, J. 2007. Soil

microbial biomass, dehydrogenase activity, bacterial community structure

in response to long-term fertilizer management. Soil Biol. Biochem., 39,

2971–2976.

Churilova, E.V. 2010. Vermicompost leachate (Vermiliquer) as a liquid fertilizer

for hydroponically-grown pak choi (Brassica chinensis L.) in the tropics.

in: Masters Thesis, Centre for Plant and Water Science, CQUniversity.

Rockhampton, Australia, pp. 215.

Clark, L.J., Whalley, W.R., Ellis-jones, J., Dent, K., Rowse, H.R., Finch-Savage,

W.E., Gatsai, T., Jasi, L., Kaseke, N.E., Murungu, F.S., Riches, C.R.,

Chiduza, C. 2001. On-farm seed priming in maize: a physiological

256
References

evaluation. Seventh Eastern and Southern Africa Regional Maize

Conference. pp. 268-273.

Clarkson, D.T. 1969. Metabolic aspects of aluminium toxicity and some possible

mechanisms for resistance. in: Ecological Aspects of Mineral Nutrition of

Plants, (Ed.) I.H. Rorison, Blackwell Sci. Publ., Oxford and Edinburgh.

Clune, T.S., Copeland, L. 1999. Effects of aluminium on canola roots. Plant Soil,

216, 27-33.

Collard, S.J., Zammit, C. 2006. Effects of land-use intensification on soil carbon

and ecosystem services in Brigalow (Acacia harpophylla) landscapes of

southeast Queensland, Australia. Agri. Ecosys. Environ., 117, 185–194.

Cook, R.J., Baker, K.F. 1983. The nature and practice of biological control of

plant pathogens. Am. Phytopathol. Soc.

Cook, R.J., Thomashow, L.S., Weller, D.M., Fujimoto, D., Mazzola, M.,

Bangera, G., Kim, D.S. 1995. Molecular mechanisms of defense by

rhizobacteria against root disease, Vol. 92, Proc. Natl. Acad. Sci. USA,

pp. 4197-4201.

Cotxarrera, L., Trillas-Gay, M.I., Steinberg, C., Alabouvette, C. 2002. Use of

sewege sludge compost and Trichoderma asperellum isolates to suppress

Fusarium wilt of tomato. Soil Biol. Biochem., 34, 467-476.

Crnko, G.S., Stall, W.M., White, J.M. 1992. Sweet corn weed control evolutions

on mineral and organic soils. pp. 326-330.

Cronin, M.J., Yohalem, D.S., Harris, R.F., Andrews, J.H. 1996. Putative

mechanism and dynamics of inhibition of the apple scab pathogen

Venturia inequalis by compost extracts. Soil Biol. Biochem., 28(9), 1241-

1249.

257
References

Dalal, R.C., Mayer, D.G. 1986. Long-term trends in fertility of soils under

continuous cultivation and cereal cropping in southern Queensland. I.

Overall changes in soil properties and trends in winter cereal yields. Aust.

J. Soil Res., 24, 265–279.

Danielle, O.P., Rai, K.S. 2006. On-farm management practices to minimise off-

site movement of pesticides from furrow irrigation. Pest Management

Sci., 62, 899–911.

Danon, M., Zmora-Nahum, S., Chen, Y., Hadar, Y. 2007. Prolonged compost

curing reduces suppression of Sclerotium rolfsii. Soil Biol. Biochem., 39,

1936-1946.

Davidson, E.A., Verchot, L.V., Henrique, C., Ackerman, I.L., Carvalho, J.E.M.

2000. Effects of soil water content on soil respiration in forests and cattle

pastures of eastern Amazonia. Biogeochem., 48, 53-69.

Dees, P.M., Ghiorse, W.C. 2001. Microbial diversity in hot synthetic compost as

revealed by PCR-amplified rRNA sequences from cultivated isolates and

extracted DNA. FEMS Microbiol. Ecol., 35(207–216).

Delhaize, E., Ryan, P.R. 1995. Aluminum toxicity and tolerance in plants. Plant

Physiol., 107, 315-321.

Deportes, I., Benoit-Guyod, J.L., Zmirou, D.E., Bouvier, M.C. 1995. Hazard to

man and environment posed by the use of urban waste compost: A

review. Sci. Total Environ., 172, 197–222.

Deportes, I., Benoit-Guyod, J.L., Zmirou, D.E., Bouvier, M.C. 1998. Microbial

disinfection capacity of municipal solid waste (MSW) composting. J.

Appl. Microbiol., 85, 238–46.

258
References

Dianez, F., Santos, M., Boix, A., Cara, M., Trillas, I., Avelis, M., Tello, J.C.

2006. Grape marc compost tea suppressiveness to plant pathogenic fungi:

Role of siderophores. Compost Sci. Util., 14(1), 48-53.

Diatloff, E., Harper, S.M., Asher, C.J., Smith, F.W. 1998. Effects of humic and

fulvic acids on the rhizotoxicity of lanthanum and aluminium to corn.

Aust. J. Soil Research, 36, 913-919.

Dilly, O., Bloem, J., Vos, A., Munch, J.C. 2004. Bacterial diversity in agricultural

soils during litter decomposition. Appl. Environ. Microbiol., 70, 468-474.

Diver, S. 2002. Notes on compost teas: A supplement to the ATTRA publication:

compost teas for plant disease control. in: Pest Management Technical

Note, Appropriate Technology Transfer for Rural Areas (ATTRA).

University of Arkansas, Fayetteville, pp. 1-19.

Döbereiner, J., Pedrosa, F.O. 1987. Nitrogen Fixing Bacteria in Non-leguminous

Crop Plants. Science Tech. Madison and Springer Verlag, Berlin.

Dominy, C.S., Haynes, R.J. 2002. Influence of agricultural land management on

organic matter content, microbial activity and aggregate stability in the

profiles of two Oxisols. Biol. Fertil. Soils, 36, 298–305.

Dominy, C.S., Haynes, R.J., van Antwerpen, R. 2002. Loss of soil organic matter

and related soil properties under long-term sugarcane production on two

contrasting soils. Biol. Fertil. Soils, 36, 350–356.

Doran, J.W., Safley, M. 1997. Defining and assessing soil health and sustainable

productivity. in: Biological Indicators of Soil Health, (Eds.) C.E.

Pankhurst, B.M. Doube, V.V.S.R. Gupta, CAB International. New York,

pp. 1–28.

259
References

Doran, J.W., Zeiss, M.R. 2000. Soil health and sustainability: managing the biotic

component of soil quality. Appl. Soil Ecol., 15, 3-11.

DPIF. 2000. Compost. in: Agriculture Notes, State of Victoria, Department of

Primary Industries. Farm Diversification Information Service, Bendigo,

pp. 1-3.

DPIPWE. 2010. Soil Organic Matter, Department of Primary Industries, Parks,

Water and Environment. Hobart, Tasmania, Australia.

Dubeikovsky, A.N., Mordukhova, E.A., Kochetkov, V.V., Polikarpova, F.Y.,

Boronin, A.M. 1993. Growth promotion of blackcurrant softwood cuttings

by recombinant strain Pseudomonas fluorescens BSP53a synthesizing an

increased amount of indole-3-acetic acid. Soil Biol. Biochem., 25, 1277–

1281.

Dumestre, A., Sauve, S., McBride, M., Baveye, P., Berthelin, J. 1999. Copper

speciation and microbial activity in long-term contaminated soils.

Archives Environ. Contamination Toxicol., 36(2), 124-131.

Dursun, A.Y., Aksu, Z. 1999. Biodegradation kinetics of Ferrous (II) cyanide

complex ions by immobilised P. fluorescens in a packed bed column

reactor. Process Biochem., 35, 615-622.

Edwards, C.A. 1998. The use of earthworms in the breakdown and management

of organic wastes. in: Earthworm Ecology, CRC Press. Boca Raton, FL,

pp. 327–354.

Edwards, C.A., Arancon, N.Q., Graytak, S. 2006. Effects of vermicompost teas

on plant growth and disease. BioCycle, 47(5), 28-31.

260
References

Egamberdiyeva, D. 2007. The effect of plant growth promoting bacteria on

growth and nutrient uptake of maize in two different soils. Appl. Soil

Ecol., 36, 184-189.

Eiland, F., Leth, M., Klamer, M., Lind, A.M., Jensen, H.E.K., Iversen, J.J.L.

2001. C and N turnover and lignocellulose degradation during composting

of miscanthus straw and liquid pig manure. Compost Sci. Util., 9(3), 186.

Ekin, Z. 2010. Performance of phosphate solubilizing bacteria for improving

growth and yield of sunflower (Helianthus annuus L.) in the presence of

phosphorus fertilizer. Afr. J. Biotechnol., 9(25), 3794-3800.

El-Masry, M.H., Khalil, A.I., Hassouna, M.S., Ibrahim, H.A.H. 2002. In situ and

in vitro suppressive effect of agricultural composts and their water

extracts on some phytopathogenic fungi. World J. Microbiol. Biotech., 18,

551-558.

Elad, Y., Malathrakis, N.E., A.J.D. 1996. Biological control of Botrytis-incited

diseases and powdery mildews in greenhouse crops. Crop Protection,

15(3), 229-240.

Elad, Y., Steinberg, D. 1994. Effect of compost water extract on grey mould

(Botrytis cinerea). Crop Protect., 13(2), 109-114.

Elad, Y., Williamson, B., Tudzynski, P., Delen, N. 2007. Botrytis: Biology,

Pathology and Control. Kluwer academic publishers, Dordrecht,

Netherlands.

Elad, Y., Yunnis, H., Katan, T. 1992. Multiple fungicide resistance to

benzimidazoles, dicarboximides and diethofencarb in field isolates of

Botrytis cinerea in Israel. Plant Pathol., 41, 41-46.

261
References

Elfstrand, S., Hedlund, K., Martensson, A. 2007. Soil enzyme activities,

microbial community composition and function after 47 years of

continuous green manuring. Appl. Soil Ecol., 35, 610–621.

Elliot, S.L., Mumford, J.D., Moraes, G.J.d. 2002. The role of resting spores in the

survival of the mite-pathogenic fungus Neozygites floridana from

Mononychellus tanajoa during dry periods in Brazil. J. Invertebrate

Pathol., 81, 148–157.

Elliott, L.F., Lynch, J.M. 1994. Biodiversity and soil resilience. in: Soil

Resilience and Sustainable Land Use, (Eds.) D.J. Greenland, I. Szabolc,

CAB International. Wallingford, UK, pp. 353–364.

Erhart, E., Burian, K., Hartl, W., Stich, K. 1999. Suppression of Pythium ultimum

by biowaste composts in relation to compost microbial biomass, activity

and content of phenolic compounds. J. Phytopathol., 147, 299–305.

Ericsson, T. 1995. Growth and shoot: root ratio of seedlings in relation to nutrient

availability. Plant Soil, 168-169, 205-214.

Etter, L. 2008. Lofty prices for fertiliser put farmers in a squeeze. Wall Street J.

Ezzi, M.I., Lynch, J.M. 2002. Cyanide catabolising enzymes in Trichoderma spp.

Enzyme Microb. Tech., 31, 1042-1047.

Fayed, T.A. 2010a. Effect of compost tea and some antioxidant applications on

leaf chemical constituents, yield and fruit quality of pomegranate. World

J. Agric. Sci., 6(4), 402-411.

Fayed, T.A. 2010b. Optimizing, yield fruit quality and nutrition status of

Roghiani olives grown in Libya using some organic extracts. J. Hort. Sci.

Ornamental Plants, 2(2), 63-78.

262
References

Feng, X., Simpson, M.J. 2009. Temperature and substrate controls on microbial

phospholipid fatty acid composition during incubation of grassland soils

contrasting in organic matter quality. Soil Biol. Biochem., 41, 804–812.

Ferreras, L., Gomez, E., Toresani, S., Firpo, I., Rotondo, R. 2006. Effect of

organic amendments on some physical, chemical and biological properties

in a horticultural soil. Bioresour. Technol., 97, 635–640.

Finlay, R.D., Rosling, A. 2006. Integrated nutrient cycles in boreal forest

ecosystems – the role of mycorrhizal fungi. in: Fungi in Biogeochemical

Cycles, (Ed.) G.M. Gadd, Published by Cambridge.

Fontvieille, D.A., Outaguerouine, A., Thevenot, D.R. 1992. Fluorescein diacetate

hydrolysis as a measure of microbial activity in aquatic systems:

application to activated sludges. Environ. Technol., 13, 531-540.

Foy, C.D. 1974. Effects of aluminium on plant growth. in: The plant root and its

environment, (Ed.) E.W. Carson, Charlottesville Univ. Press. Virginia.

Foy, C.D. 1992. Soil chemical factors limiting plant root growth. Adv. Soil Sci.,

19, 97–149.

Foy, C.D., Chaney, R.L., M.C., W. 1978. The physiology of metal toxicity in

plants. Annu. Rev. Plant Physiol., 29, 511-566.

Franche, C., Lindström, K., Elmerich, C. 2009. Nitrogen-fixing bacteria

associated with leguminous and non-leguminous plants. Plant Soil, 321,

35–59.

Freeman, G.G., Baillie, A.J., Macinnes, C.A. 1948. Bacterial degradation of

CMC and methyl ethyl cellulose. Chem. & Industry, 279-282.

263
References

Garcia-Gil, J.C., Plaza, C., Soler-Rovira, P., Polo, A. 2000. Long-term effects of

municipal solid waste compost application on soil enzyme activities and

microbial biomass. Soil Biol. Biochem., 32, 1907-1913.

Garcıa-Gomez, A., Roig, A., Bernal, M.P. 2003. Composting of the solid fraction

of olive mill wastewater with olive leaves: organic matter degradation and

biological activity. Bioresour. Technol., 86, 59–64.

Garcia-Ochoa, F., Gomez, E., Santos, V.E., Merchuk, J.C. 2010. Oxygen uptake

rate in microbial processes: An overview. Biochem. Eng. J., 49, 289–307.

Garcia, C., Hernandez, T., Costa, F. 1991. Study on water extract of sewage

sludge composts. Soil Sci, Plant Nutr., 37(3), 399-408.

Garcia, M.I., Cruz Sosa, F., Saavedra, A.L., Hernandez, M.S. 2002. Extraction of

auxin-like substances from compost. Crop Research, 24, 323-327.

Garg, S.K., Bhatnagar, A., Kalla, A., Narula, N. 2001. In vitro nitrogen fixation,

phosphate solubilization, survival and nutrient release by Azotobacter

strains in an acquatic systems. Bioresour. Technol., 80, 101-109.

Garland, J.L. 1997. Analysis and interpretation of community-level physiological

profiles in microbial ecology. FEMS Microbiol. Ecol., 24, 289-300.

Garland, J.L. 1996. Analytical approaches to the characterization of samples of

microbial communities using patterns of potential C source utilization.

Soil Biol. Biochem., 28, 213–221.

Garland, J.L. 1999. Potential and limitations of BIOLOG for microbial

community analysis. Microbial Biosystems: New Frontiers: Proceedings

of the 8th International Symposium on Microbial Ecology, Halifax,

Canada. Atlantic Canada society for microbial ecology.

264
References

Garland, J.L., Mills, A.L. 1991. Classification and characterization of

heterotrophic microbial communities on the basis of patterns on

community-level, sole-carbon-source utilization. Appl. Environ.

Microbial., 57, 2351–2359.

GENSTAT. 2008. GenStat Procedure Library Release PL16. Eleventh ed.

Gerhardt, R.A. 1997. A comparative analysis of the effects of organic and

conventional farming systems on soil structure. Biol. Agricultur.

Horticol., 14, 139-157.

Giller, K.E., Beare, M.H., Lavelle, P., Izac, A.M.N., Swift, M.J. 1997.

Agricultural intensification, soil biodiversity and agroecosystem function.

Appl. Soil Ecol., 6, 3–16.

Girvan, M.S., Bullimore, J., Ball, A.S., Pretty, J.N., Osborn, A.M. 2004.

Responses of active bacterial and fungal communities in soils under

winter wheat to different fertilizer and pesticide regimens. Appl. Environ.

Microbiol., 70, 2692–2701.

Girvan, M.S., Bullimore, J.N., Osborn, A.M., Ball, A.S. 2003. Soil type is the

primary determinant of the composition of the total and active bacterial

communities in arable soil. Appl. Environ. Microbiol., 69, 1800-1809.

Golchin, A., Clarke, P., Oades, J.M., Skjemstad, J.O. 1995. The effects of

cultivation on the composition of organic matter and structural stability of

soils. Aust. J. Soil Res., 33, 975–993.

Gomez, E., Ferreras, L., Toresani, S., Ausilio, A., Bisaro, V. 2001. Changes in

some soil properties in a vertic argiudoll under shortterm conservation

tillage. Soil Till. Res., 61, 179–186.

265
References

Graham, M.H., Haynes, R.J. 2005. Catabolic diversity of soil microbial

communities under sugarcane and other land uses estimated by Biolog

and substrate-induced respiration methods. Appl. Soil Ecol., 29, 155–164.

Graham, M.H., Haynes, R.J., Meyer, J.H. 2002. Soil organic matter content and

quality: Effects of fertilizer applications, burning and trash retention on a

long-term sugarcane experiment in South Africa. Soil Biol. Biochem., 34,

93-102.

Grayston, S.J., Campbell, C.D., Bardgett, R.D., Mawdsley, J.L., Clegg, C.D.,

Ritz, K., Griffiths, B.S., Rodwell, J.S., Edwards, S.J., Davies, W.J.,

Elston, D.J., Millard, P. 2004. Assessing shifts in microbial community

structure across a range of grasslands of differing management intensity

using CLPP, PLFA and community DNA techniques. Appl. Soil Ecol., 25,

63–84.

Greathouse, G. 1949. Microbiological degradation of cellulose. Am. Chem.Soc.

Meeting.

Green, V.S., Stott, D.E., Diack, M. 2006. Assay for fluorescein diacetate

hydrolytic activity: Optimisation for soil samples. Soil Biol. Biochem., 38,

693-701.

Griffiths, E. 1965. Micro-organisms and soil structure. Biol. Rev., 40(1), 129–

142.

Grobe, K. 2003. California landscape contractor calls it compost tea time.

BioCycle, 44(2), 26-27.

Grundy, A.C., Green, J.M., Lennartsson, M. 1998. The effect of temperature on

the viability of weed seeds in compost. Compost Sci. Util., 6, 26–33.

266
References

Gurbuz, F., Ciftci, H., Akcil, A. 2009. Biodegradation of cyanide containing

effluents by Scenedesmus obliquus. J. Hazard. Mater., 162, 74-79.

Ha, K.V., Marschner, P., Bünemann, E.K. 2008. Dynamics of C, N, P and

microbial community composition in particulate soil organic matter

during residue decomposition. Plant Soil 303, 253–264.

Haaba, D., Hagspiela, K., Szakmarya, K., Kubicek, C.P. 1990. Formation of the

extracellular proteases from Trichoderma reesei QM 9414 involved in

cellulase degradation. J. Biotechnol., 16(3-4), 187-198.

Hadar, Y., Gorodecki, B. 1991. Suppression of germination of sclerotia of

Sclerotium rolfsii in compost. Soil Biol. Biochem., 23, 303-306.

Haggag, W.M., Saber, M.S.M. 2007. Suppression of early blight on tomato and

purple blight on onion by foliar sprays of aerated and non-aerated

compost teas. J. Food Agri. Environ., 5(2), 302-309.

Hall, S.G., Schellinger, D.A., Carney, W.A. 2006. Enhancing sugarcane field

residue biodegradation by grinding and use of compost tea. Compost Sci.

Util., 14, 32-38.

Halsall, D.M., Gibson, A.H. 1985. Cellulose decomposition and associated

nitrogen fixation by mixed cultures of Cellulomonas gelida and

Azospirillum Species or Bacillus macerans. Appl. Environ. Microbiol.,

50(4), 1021-1026.

Han, G.X., Zhou, G.S., Xu, Z.Z., Yang, Y., Liu, J.L., Shi, K.Q. 2007. Soil

temperature and biotic factors drive the seasonal variation of soil

respiration in a maize (Zea mays L.) agricultural ecosystem. Plant Soil,

291, 15–26.

267
References

Hardy, G.E., Sivasithamparam, K. 1991. Effects of sterile and non-sterile

leachates extract from composted Eucalyptus-bark and Pine-bark

container on Phytophthora spp. Soil Biol. Biochem., 23, 25-30.

Hargreaves, J.C., Adl, M.S., Warman, P.R. 2009. The effects of municipal solid

waste compost and compost tea on mineral element uptake and fruit

quality of strawberries. Compost Sci. Util., 17(2), 85-94.

Hargreaves, J.C., Adl, M.S., Warman, P.R., Rupasinghe, H.P.V. 2008. The

effects of organic amendments on mineral element uptake and fruit

quality of raspberries. Plant Soil, 308, 213-226.

Harman, G.E. 2006. Overview of mechanisms and uses of Trichoderma spp.

Phytopathol., 96(2), 190-194.

Harper, S.M., Edwards, D.G., Kerven, G.L., Asher, C.J. 1995. Effects of organic

acid fractions extracted from Eucalyptus camaldulensis leaves on root

elongation of maize (Zea mays) in the presence and absence of

aluminium. Plant Soil, 171, 189-192.

Harrower, K.M. 2006. Assessing fungal growth in the undergraduate teaching

laboratory. Aust. Mycol., 25(1), 1-4.

Hattenschwiler, S., Tiunov, A.V., Scheu, S. 2005. Biodiversity and litter

decomposition in terrestrial ecosystems. Annu. Rev. Ecol. Evol. Syst., 36,

191–218.

Heal, O.W., Anderson, J.W., Swift, M.J. 1997. Plant litter quality and

decomposition: an historical overview. in: Driven by Nature: Plant Litter

Quality and Decomposition, (Eds.) G. Cadisch, K.E. Giller, CAB

International. Wallingford, UK, pp. 3-30.

268
References

Henis, Y., Lewis, J.A., Papavizas, G.C. 1984. Interactions between Sclerotium

rolfsii and Trichoderma spp. - relationship between antagonism and

disease control. Soil Biol. & Biochem., 16, 391-395.

Henning, S.J. 1975. Aluminium toxicity in the primary meristem of wheat roots,

Vol. PhD, PhD thesis, Oregon State University. Oregon, United States.

Hernandez, D., Fernandez, J.M., Plaza, C., Polo, A. 2007. Water-soluble organic

matter and biological activity of a degraded soil amended with pig slurry.

Sci. Tot. Environ., 378, 101-103.

Herrmann, R.F., Shann, J.F. 1997. Microbial community changes during the

composting of municipal solid waste. Microb. Ecol., 33, 78–85.

Hibar, K., Daami-Remadi, M., Jabnoun-Khiareddine, H., Znaïdi, I.E., El-

Mahjoub, M. 2006. Effect of compost tea on mycelial growth and disease

severity of Fusarium oxysporum f. sp. radicis-lycopersici. Biotechnol.

Agron. Soc. Environ., 10(2), 101-108.

Highley, T.L. 1980. Cellulose degradation by cellulose clearing and non-cellulose

clearing brown rot fungi. Appl. Environ. Microbiol., 40(6), 1145-1147.

Hoitink, H.A.J., Boehm, M.J. 1999a. Biocontrol within the context of soil

microbial communities: A substrate-dependent phenomenon. Annu. Rev.

Phytopathol., 37, 427–446.

Hoitink, H.A.J., Boehm, M.J. 1999b. Biocontrol within the context of soil

microbial communities: A substrate-dependent phenomenon. Annu.

Rev.Phytopathol., 37, 427–446.

Hoitink, H.A.J., Stone, A.G., Han, D.Y. 1997. Suppression of plant diseases by

composts. Hort. Sci., 32, 184-187.

269
References

Holtan-Hartwig, L., Dorsch, P., Bakken, L.R. 2002. Low temperature control of

soil denitrifying communities: kinetics of N2O production and reduction.

Soil Biol. Biochem., 34, 1797–1806.

Hoyos-Carvajal, L., Orduz, S., Bissett, J. 2009. Growth stimulation in bean

(Phaseolus vulgaris L.) by Trichoderma. Biol. Cont., 51, 409–416.

Hu, W.H., Zhou, Y.H., Du, Y.S., Xia, X.J., Yu, J.Q. 2006. Differential response

of photosynthesis in greenhouse- and field-ecotypes of tomato to long-

term chilling under low light. J. Plant Physiol., 163, 1238—1246.

Hue, N.V., Amien, I. 1989. Aluminium detoxification with green manures.

Commun. Soil Sci. Plant Anal., 20(15 & 16), 1499-1511.

Imray, P., Moore, M.R., Callan, P.W., Lock, W. 1995. Aluminium.

Ingham, E.R. 1998. Anaerobic bacteria and compost tea. BioCycle, 39(6), 86-86.

Ingham, E.R. 2000. Brewing Compost Tea. Kitchen Gardener, 29.

Ingham, E.R. 2003. Compost tea – promises & practicalities. in: Reprinted from

ACRESUSA, a voice for eco-agriculture, Vol. 33.

Ingham, E.R. 2005. The Compost Tea Brewing Manual. Fifth ed. US Printings,

Soil Foodweb Incorporated, Oregon.

Ingham, E.R. 2004. Compost Tea Quality: Light Microscope Methods. Soil

Foodweb Inc, Corvallis, Oregon

Ingham, E.R. 1999a. Making a high quality compost tea, Part II. BioCycle, 40(4),

94.

Ingham, E.R. 1999b. What is compost tea? Part I. BioCycle, 40(3), 74-75.

Ingham, E.R., Alms, M. 1999. Compost Tea Manual. Soil Food Web, Growing

Solutions Inc., Corvallis, OR.

270
References

Insam, H. 1997. Substrate utilisation test in microbial ecology: A preface to the

special issue. J. Microbial Methods, 30, 1-2.

Jackson, W.A. 1967. Physiological effects of soil acidity. in: Soil Acidity and

Liming, (Eds.) R.W. Pearson, F. Adams, Vol. 12, Am. Soc. agron.

Madison, Wis, pp. 43-124.

Janvier, C., Villeneuve, F., Alabouvette, C., Edel-Hermann, V., Mateille, T.,

Steinberg, C. 2007. Soil health through soil disease suppression: Which

strategy from descriptors to indicators? Soil Biol. Biochem., 39, 1–23.

Jany, J., Barbier, G. 2008. Culture-independent methods for identifying microbial

communities in cheese. Food Microbiol., 25, 839-848.

Janzen, R.A., Cook, F.D., McGill, W.B. 1995. Compost extract added to

microcosms may stimulate community-level controls on soil

microorganisms involved in element cycling. Soil Biol. Biochem., 27(2),

181-188.

Jesus, E.d.C., Marsh, T.L., Tiedje, J.M., Moreira, F.M.d.S. 2009. Changes in land

use alter the structure of bacterial communities in Western Amazon soils.

Int. Soc. Micro. Ecol. J., 3, 1004–1011.

Jing, H., Liu, H., Wong, T., Chen, M. 2010. Impact of mesozooplankton grazing

on the microbial community revealed by denaturing gradient gel

electrophoresis (DGGE). J. Experimental Marine Biol. Ecol., 383, 39–47.

Jones, D., Wyatt, D. 2005. Super Brewing Guidelines: The Vermicrobe System

Manual.

Jones, K., Bangs, D. 1985. Nitrogen fixation by free-living heterotrophic bacteria

in an oak forest: The effect of liming. Soil Biol. Biochem., 17(5), 705-709.

271
References

Jones, P., Martin, M. 2003. A review of the literature on the occurance and

survival of pathogens of animals and humans in green compost. Institute

for Animal Health, Compton.

Joshi, D., Hooda, K.S., Bhatt, J.C., Mina, B.L., Gupta, H.S. 2009. Suppressive

effects of composts on soil-borne and foliar diseases of French bean in the

field in the western Indian Himalayas. Crop Protec., 28, 608–615.

Kahindi, J.H.P., Woomer, P., George, T., Moreira, F.M.d.S., Karanja, N.K.,

Giller, K.E. 1997. Agricultural intensification, soil biodiversity and

ecosystem function in the tropics: the role of nitrogen-fixing bacteria.

Appl. Soil Ecol., 6, 55-76.

Kai, H., Udea, T., Sakaguchi, M. 1990. Antimicrobial activity of bark compost

extracts. Soil Biol. Biochem., 7, 983-986.

Kaiser, E.-A., Martens, R., Heinemeyer, O. 1995. Temporal changes in soil

microbial biomass carbon in an arable soil:consequences for soil

sampling. Plant Soil, 170, 287-295.

Kale, R.D., Mallesh, B.C., Bano, K., Bagyaraj, D.J. 1992. Influence of

vermicompost applications on the available macro nutrients and selected

microbial population in a paddy field. Soil Biol. Biochem., 24(12), 1317-

1320.

Kelley, S. 2004. Building a knowledge base for compost tea. BioCycle, 45(6), 30-

34.

Ketterer, N., Fisher, B., Weltzien, H.C. 1992. Biological control of Botrytis

cinerea on grapevine by compost extracts and their microorganisms in

pure culture. Proceedings of the 10th International Botrytis Symposium,

Wageningen, The Netherlands. pp. 179-186.

272
References

Keyser, H.H., Munns, D.N. 1978. Tolerance of rhizobia to acidity, aluminum and

phosphate. Soil Sci. Soc.Am. J., 43, 519-523.

Khammas, K.M., Kaiser, P. 1992. Pectin decomposition and associated nitrogen

fixation by mixed cultures of Azospirillum and Bacillus species. Can. J.

Microbiol., 38, 794-797.

Kingston, G., Norris, C. 2001. The green cane harvesting system - an Australian

perspective. in: Innovative approaches to sugarcane productivity in the

new millennium, American Society of Sugar Cane Technologists. Miami,

Florida, pp. 9.

Kloepper, J.W., Zehnder, G.W., Tuzun, S., Murphy, J.F., Wei, G., Yao, C.,

Raupach, G.S. 1996. Toward agricultural implementation of PGPR-

mediated induced systemic resistance against crop pests. in: Advances in

Biological Control of Plant Diseases, (Eds.) T. Wenhua, R.J. Cook, A.

Rovira, China Agricultural University Press. Beijing, pp. 165-174.

Koné, S.B., Dionne, A., Tweddell, R.J., Antoun, H., Avis, T.J. 2010. Suppressive

effect of non-aerated compost teas on foliar fungal pathogens of tomato.

Biol. Cont., 52, 167–173.

Konopka, A., Oliver, L., Turco, J. 1998. The use of carbon substrate utilization

patterns in environmental and ecological microbiology. Microb. Ecol., 35,

103-115.

Koschinsky, S., Peters, S., Schwieger, F., Tebbe, C.C. 1999. Applying molecular

techniques to monitor microbial communities in composting processes. in:

Microbial Biosystems: New Frontiers, Microbial Processes during

Composting, (Eds.) C.R. Bell, M. Brylinsky, P. Johnson-Green. Atlantic

Canada Society for Microbial Ecology, Halifax, Canada.

273
References

Krebs, C.J. 1999. Ecological Methodology, (Ed.) B. Cummings. Menlo Park,

California, pp. 620.

Lalande, R., Gagnon, B., Simard, R.R., Cote, D. 2000. Soil microbial biomass

and enzyme activity following liquid hog manure application in a long-

term field trial. Can. J. Soil Sci., 80, 263-269.

Lampkin, N. 1990. Organic Farming. in: Farming Press Books, Ipswich, UK.

Larkin, R.P. 2008. Relative effects of biological amendments and crop rotations

on soil microbial communities and soilborne diseases of potato. Soil Biol.

Biochem., 40, 1341-1351.

Lauber, C.L., Strickland, M.S., Bradford, M.A., Fierer, N. 2008. The influence of

soil properties on the structure of bacterial and fungal communities across

land-use types. Soil Biol. Biochem., 40, 2407–2415.

Lee, K.E., Pankhurst, C.E. 1992. Soil organisms and sustainable productivity.

Aust. J. Soil Research, 30(6), 855-892.

Lemenih, M., Karltun, E., Olsson, M. 2005. Assessing soil chemical and physical

property responses to deforestation and subsequent cultivation in

smallholders farming system in Ethiopia. Agric. Ecosyst. Environ., 105,

373–386.

Leu, A. 2006. Organics and soil carbon: increasing soil carbon, crop productivity

and farm profitability. Third OFA National Organic Conference,

“Organics – Solutions to Climate Change”, Sydney, Australia. pp. 4-12.

Liang, Y., Nikolic, M., Peng, Y., Chen, W., Jiang, Y. 2005. Organic manure

stimulates biological activity and barley growth in soil subject to

secondary salinization. Soil Biol. Biochem., 37, 1185-1195.

274
References

Linehan, D.J. 1976. Some effects of a fulvic acid component of soil organic

matter on the growth of cultured excised tomato roots. Soil Biol.

Biochem., 8, 511-517.

Lines-Kelly, R. 2004. Current research into soil biology in agriculture. NSW.

Department of Primary Industries.

Litterick, A.M., Harrier, L., Wallace, P., Watson, C.A., Wood, M. 2004. The role

of uncomposted materials, composts, manures, and compost extracts in

reducing pest and disease incidence and severity in sustainable temperate

agricultural and horticultural crop production - A review. Critical Rev.

Plant Sci., 26(3), 453-479.

Liu, B., Tu, C., Hu, S., Gumpertz, M., Ristaino, J.B. 2007. Effect of organic,

sustainable, and conventional management strategies in grower fields on

soil physical, chemical, and biological factors and the incidence of

Southern blight. Appl. Soil Ecol., 37, 204-214.

Liu, L., Kloepper, J.W., Tuzun, S. 1995. Induction of systemic resistance in

cucumber against bacterial angular leaf spot by plant growth-promoting

rhizobacteria. Phytopathol., 85, 843-847.

Lloyd, J., Taylor, J.A. 1994. On the temperature dependence of soil respiration.

Functional Ecol., 8, 315–323.

Ma, J.F., Ryan, P.R., Delhaize, E. 2001. Aluminium tolerance in plants and the

complexing role of organic acids. Trends Plant Sci., 6(6), 273-278.

Ma, L., Xiongwu, Q., Fen, G., Bianqing, H. 2001b. Control of sweet pepper

Fusarium wilt extracts and compost with its mechanism. Chin. J. Appl.

Env. Biol., 7(1), 84–87.

275
References

Mabbutt, J.A. 1992. Degradation of the Australian drylands: a historical

approach. in: Degradation and Restoration of Arid Lands, (Ed.) H.E.

Dregne, pp. 27–98.

Maggioni, A., Varanini, Z., Nardi, S., Pinton, R. 1987. Action of soil humic

matter on plant roots: stimulation of ion uptake and effects on (Mg 2+ + K

+) ATPase activity. Sci. Total Environ., 62, 355-363.

Magid, J., De Neergaard, A., Brandt, M. 2006. Heterogenous distribution may

substantially decrease initial decomposition, long-term microbial growth

and N-immobilization from high C-to-N ratio resources. Eur. J. Soil Sci.,

57(517-529).

Mann, L.K. 1986. Changes in soil carbon storage after cultivation. Soil Sci., 142,

279–288.

Marcote, I., Hernandez, T., Garcia, C., Polo, A. 2001. Influence of one or two

successive annual applications of organic fertilizers on the enzyme

activity of a soil under barley cultivation. Bioresour. Technol., 79(2), 147-

154.

Marmonier, P., Fontvieille, D., Gibert, J., Vanek, V. 1995. Distribution of

dissolved organic carbon and bacteria at the interface between the Rhone

River and its alluvial aquifer. J. North Am. Benthol. Soc., 14, 382-392.

Martens, M.-H.R. 2001. Compost tea - Just what the doctor ordered. A Voice for

Eco-Agriculture, 31(2), 13.

Mato, M.C., Gonzalez-Alonso, L.M., Mendez, J. 1972a. Inhibition of enzymatic

indoleacetic acid oxidation by soil fulvic acids. Soil Biol. Biochem., 475-

478.

276
References

Mato, M.C., Olmedo, M.G., Mendez, J. 1972b. Inhibition of indoleacetic acid-

oxidase by soil humic acids fractionated on sephadex. Soil Biol. Biochem.,

4, 469-473.

Matsumoto, H., Hirasawa, R., Toricari, H., Takahashi, E. 1976. Localisation of

absorbed aluminium in pea roots and its binding to nucleic acids. Plant

Cell Physiol., 18, 325-335.

Matsumoto, H., Morimura, S., Takahashi, E. 1977. Less environment of pectin in

the precipitation of aluminium in pea roots. Plant Cell Physiol., 17(127-

137).

Mazzola, M. 2004. Assessment and management of soil microbial community

structure for disease suppression. Annu. Rev. Phytopathol., 42, 35-59.

McCaig, A.E., Grayston, S.J., Prosser, J.I., Glover, L.A. 2001. Impact of

cultivation on characterisation of species composition of soil bacterial

communities. FEMS Microbiol. Ecol., 35, 37-48.

McQuilken, M.P., Whipps, J.M., Lynch, J.M. 1994. Effects of water extracts of a

composted manure-straw mixture on the plant pathogen Botrytis cinerea.

World J. Microbiol. Biotech., 10, 20-26.

Meier, E. 2007. Availability of nitrogen in green cane trash blanketed soils in the

wet tropics and its impact on productivity / profitability: a systems

analysis. in: School of Natural and Rural Systems Management, Vol. PhD,

The University of Queensland. Brisbane, pp. 231.

Melero, S., Porras, J.C.R., Herencia, J.F., Madejon, E. 2006. Chemical and

biochemical properties in a silty loam soil under conventional and organic

management. Soil Tillage Res., 90, 162–170.

277
References

Merrill, R., Hoberechi, K., McKeon, J. 1997. Organic teas from composts and

manures. Cabrillo Community College.

Metcalf, Eddy, I. 2003. Wastewater Engineering: Treatment and Reuse. Fourth

ed. McGraw-Hill International editions, New York.

Michel, F.C., Forney, L.J., Huang, A.J.F., Drew, S., Czuprenski, M., Lindeberg,

J.D., Reddy, C.A. 1996a. Effects of turning frequency, leaves to grass mix

ratio and windrow vs. pile configuration on the composting of yard

trimming. Compost Sci. Util., 4(1), 26-43.

Michel, F.C., Huang, A.J.F., Forney, L.J., Reddy, C.A. 1996b. Field scale study

of the effect of pile size, turning regime and leaf to grass mix ratio on the

composting of yard trimmings. in: The Science of Composting Part 1,

(Eds.) M.D. Bertoldi, P. Sequi, P. Lemmes, T. Papi. London, U.K., pp.

577-584.

Midmore, D.J. 2006. Intensive organic production systems and issues of

sustainability. Primary Industries Research Centre, Central Queensland

University, Australia.

Midmore, D.J., Wu, D.L. 1999. Work that water! Hydroponics made easy.

Waterlines, 17(4), 28-30.

Misra, R.V., Roy, R.N., Hiraoka, H. 2003. On-farm composting methods.

Mitchell, R.D.J., Thorburn, P.J., Larsen, P. 2000. Quantifying the loss of

nutrients from the immediate area when sugarcane residues are burnt.

Proceedings of the Australian Society of Sugar Cane Technologists. pp.

206-211.

Moeskops, B., Sukristiyonubowo, Buchan, D., Sleutel, S., Herawaty, L., Husen,

E., Saraswati, R., Setyorini, D., De Neve, S. 2010. Soil microbial

278
References

communities and activities under intensive organic and conventional

vegetable farming in West Java, Indonesia. Appl. Soil Ecol., 45, 112-120.

Moon, P. 1997. Final report on basic on-farm composting manual. CM-97-3.

Munns, D.N. 1977. Soil acidity and related factors. in: Exploiting the Legume-

Rhizobium Symbiosis in Tropical Agriculture, (Eds.) J.M. Vincent, A.J.

Whitney, J.B. Niftal. Hawaii, pp. 211-236.

Munns, D.N., Keyser, H.H. 1981. Response of Rhizobium strains to acid and

aluminium stress. Soil Biol. Biochem., 13, 115-118.

Murray, A.E., Hollibaugh, J.T., Orrego, C. 1996. Phylogenetic compositions of

bacterioplankton from two Californian estuaries by denaturing gradient

gel electrophoresis of 16S rDNA fragments. App. Environ. Microbiol., 59,

2676-2680.

Muscolo, A., Bovalo, F., Gionfriddo, F., Nardi, S. 1999. Earthworm humic matter

produces auxin-like effects on Daucus carota cell growth and nitrate

metabolism. Soil Biol. Biochem., 31, 1303-1311.

Mussaddak, J., Fawaz, K. 2008. Assessment of injection of liquid rhizobial

inoculum and traditional inoculation of soybean under furrow and drip

irrigation. Agrochimica, 52(1), 1-11.

Muyzer, G. 1999. DGGE/TGGE a method for identifying genes from natural

ecosystems. Curr. Opin. Microbiol., 2(3), 317-322.

Muyzer, G., Smalla, K. 1998. Application of denaturing gradient gel

electrophoresis (DGGE) and temperature gradient gel electrophoresis

(TGGE) in microbial ecology. Antonie Van Leeuwenhoek, 73(1), 127-41.

Muyzer, G., Wall, E.C.D., Uitterlinden, A.G. 1993. Profiling of complex

microbial populations by denaturing gradient gel electrophoresis analysis

279
References

of polymerase chain reaction-amplified genes coding for 16S rRNA. Appl.

Environ. Microbiol., 59, 695–700.

Naidu, Y., Meon, S., Kadir, J., Siddiqui, Y. 2010. Microbial starter for the

enhancement of biological activity of compost tea. Int. J. Agric. Biol., 12,

51–56.

Ndegwa, P.M., Thompson, S.A. 2000. Effects if C-to-N ratio on

vermicomposting of biosolids. Bioresour. Technol., 75, 7-12.

Nehls, U. 2008. Mastering ectomycorrhizal symbiosis: the impact of

carbohydrates. J. Experimental Bot., 59(5), 1097–1108.

Nelson, E.B., Hoitink, H.A.J. 1983. The role of microorganisms in the

suppression of Rhizoctonia solani in container media amended with

composted hardwood bark. Phytopathol., 73, 274-278.

Nelson, P.N., Oades, J.M. 1998. Organic matter, sodicity, and soil structure. in:

Sodic Soils, (Eds.) M.E. Summer, R. Naidu, Oxford University Press.

New York, pp. 51-75.

Neuhauser, E.F., Loehr, R.C., Malecki, M.R. 1988. The potential of earthworms

for managing sewage sludge. in: Earthworms in Waste and Environmental

Management, (Eds.) C.A. Edwards, E.F. Neuhauser, SPB Academic

Publishing. Netherlands, pp. 9–20.

Nilsson, T. 1974. Comparative study on the cellulolytic activity of white-rot and

brown-rot fungi. Mater. Org., 9, 173-198.

Nilsson, T., Ginns, J. 1979. Cellulolytic activity and the taxonomic position of

selected brown-rot fungi. Mycologia, 71, 170-177.

280
References

Norrish, S. 1996. Constraints to the adoption of green cane trash blanketing in

central and southern districts. Sugar Research and Development

Corporation. BS109S.

NOSB. 2004. Compost tea task force report. National Organic Standards Board.

Nosko, P., Brassard, P., Kramer, J.R., Kershaw, K.A. 1988. The effect of

aluminium on seed germination and early seedling establishment growth

and respiration of white spruce (Picea glauca). Can. J. Bot., 66, 2305-

2310.

Ntougias, S., Ehaliotis, C., Papadopoulou, K.K., Zervakis, G. 2006. Application

of respiration and FDA hydrolysis measurements for estimating microbial

activity during composting processes. Biol. Fertility Soil, 42, 330-337.

Ntougias, S., Zervakis, G.I., Kavroulakis, N., Ehaliotis, C., Papadopoulou, K.K.

2004. Bacterial diversity in spent mushroom compost assessed by

amplified rDNA restriction analysis and sequencing of cultivated isolates.

Systematic Appl. Microbiol., 27, 746-754.

Nusslein, K., Tiedje, J.M. 1999. Soil bacterial community shift correlated with

change from forest to pasture vegetation in a tropical soil. Appl. Environ.

Microbiol., 65, 3622-3626.

O‘Grady, R., Alexander, M. 2008. Healthy Soils Workshop. Australia.

Okur, N. 2002. Response of soil biological and biochemical activity to salination.

The Journal of Agricultural Faculty of Ege University, 39(1), 87-93.

Olanya, O.M., Larkin, R.P. 2006. Efficacy of essential soils and biopesticides on

Phytopthora infestans suppression in laboratory and growth chamber

studies. Biocontrol Sci.Technol., 16(9), 901-917.

281
References

Oldeman, L., van Engelen, V., Pulles, J. 1990. The Extent of Human-induced Soil

Degradation. International Soil Reference and Information Centre,

Wageningen.

Omay, A.B., Rice, C.W., Maddux, L.D., Gordon, W.B. 1998. Corn yield and

nitrogen uptake in monoculture and in rotation with soybean. Soil Sci.

Soc. Am. J., 62, 1596-1603.

OSUE. 1999. Ohio’s Livestock and Poultry Mortality Composting Manual. Ohio

State Univ. Ext. Pub.

Ovreas, L., Jensen, S., Daae, F.L., Torsvik, V. 1998. Microbial community

changes in a perturbed agricultural soil investigated by molecular and

physiological approaches. Appl. Environ. Microbiol., 64, 2739-2742.

Pankhurst, C.E., Doube, B.M., Gupta, V.V.S.R. 1997. Biological indicators of

soil health: synthesis. in: Biological indicators of soil health, (Eds.) C.E.

Pankhurst, B.M. Doube, V.V.S.R. Gupta, CAB International. New York,

pp. 419–435.

Pant, A.P., Radovich, T.J.K., Hue, N.V., Talcott, S.T., Krenek, K.A. 2009.

Vermicompost extracts influence growth, mineral nutrients,

phytonutrients and antioxidant activity in pak choi (Brassica rapa cv.

Bonsai, Chinensis group) grown under vermicompost and chemical

fertiliser. J. Sci. Food Agri.

Parkin, T.B., Doran, J.W., Franco-Vizcaino, E. 1996. Field and laboratory tests of

soil respiration. in: Methods for Assessing Soil Quality, (Eds.) J.W. Doran,

A.J. Jones, SSSA Special Publication, pp. 231–245.

Parmelee, R.W., Beare, M.H., Cheng, W., Hendrix, P.F., Rider, S.J., Crossley,

D.A., Coleman, D.C. 1990. Earthworms and enchytraeids in conventional

282
References

and no tillage agroecosystems: a biocide approach to assess their role in

organic matter breakdown. Biol. Fertility Soils, 10, 1-10.

Parsons, S.A., Congdon, R.A. 2008. Plant litter decomposition and nutrient

cycling in north Queensland tropical rain-forest communities of differing

successional status. J. Tropical Ecol., 24, 317–327.

Pascual, J.A., Hernandez, T., Garcia, C., Ayuso, M. 1998. Enzymatic activities in

an arid soil amended with urban organic wastes: laboratory experiment.

Bioresour. Technol., 64, 131-138.

Pe´rez-Piqueres, A., Edel-Hermann, V., Alabouvette, C., Steinberg, C. 2006.

Response of soil microbial communities to compost amendments. Soil

Biol. Biochem., 38, 460–470.

Pe´rez, J., Mun˜oz-Dorado, J., de la Rubia, T., Martı´nez, J. 2002. Biodegradation

and biological treatments of cellulose, hemicellulose and lignin: an

overview. Int. Microbiol., 5, 53–63.

Peacock, A.D., Mullen, M.D., Ringelberg, D.B., Tyler, D.D., Hedrick, D.B.,

Gale, P.M., White, D.C. 2001. Soil microbial community responses to

dairy manure or ammonium nitrate applications. Soil Biol. Biochem., 33,

1011-1019.

Phae, C.G., Sasaki, M., Shoda, M., Kubota, H. 1990. Characteristics of Bacillus

subtilis isolated from composts suppressing phytopathogenic

microorganisms. Soil Sci. Plant Nutr., 36, 575-586.

Piccolo, A., Nardi, S., Concheri, G. 1992. Structural characteristics of humic

substances as related to nitrate uptake and growth regulation in plant

systems. Soil. Biol. Biochem., 24, 373-380.

283
References

Postma, J., Schilder, M.T., Bloem, J., van Leeuwen-Haagsma, W.K. 2008. Soil

suppressiveness and functional diversity of the soil microflora in organic

farming systems. Soil Biol. Biochem., 40, 2394-2406.

Potera, C. 1994. From bacteria: A new weapon against fungal infection. Science,

265, 605.

Prasanna, R., Jaiswal, P., Singh, Y.V., Singh, P.K. 2008. Influence of

biofertilizers and organic amendments on nitrogenase activity and

phototrophic biomass of soil under wheat. Acta Agron. Hung, 56(2), 149-

159.

Preston-Mafham, J., Boddy, L., Randerson, P.F. 2002. Analysis of microbial

community functional diversity using sole-carbon-source utilisation

profiles – a critique. FEMS Microbiol. Ecol., 42, 1-14.

Pulleman, M., Jongmans, A., Marinissen, J., Bouma, J. 2003. Effects of organic

versus conventional arable farming on soil structure and organic matter

dynamics in a marine loam in the Netherlands. Soil Use and Management,

19(2), 157-165.

Radford, B.J., Thornton, C.M., Cowie, B.A., Stephens, M.L. 2007. The Brigalow

Catchment Study: III. Productivity changes on brigalow land cleared for

long-term cropping and for grazing. Australian Journal of Soil Research,

45, 512–523.

Raghothama, K.G. 1999. Phosphate acquisition. Ann. Rev. Plant Physiol.

Molecular Biol., 50, 665–693.

Ramsay, M.D., O‘Brien, R.G., Pegg, K.G. 1992. Fusarium oxysporum associated

with wilt and root rot of tomato in Queensland; races and vegetative

compatibility groups. Aust. J. Exp. Agri., 32, 651-655.

284
References

Rao, D.L.N., Pathak, H. 1996. Ameliorative influence of organic matter in the

biological activity of salt affected soils. Arid Soil Res. Rehab., 10, 311-

319.

Rashid, A., Harris, D., Hollington, P., Ali, S. 2004. On-farm seed priming

reduces yield losses of mungbean (Vigna radiata) associated with

mungbean yellow mosaic virus in the North West Frontier Province of

Pakistan. Crop Protection, 23, 1119–1124.

Raupach, G.S., Kloepper, J.W. 1997. Integrated pest management of multiple

cucumber pathogens through PGPR-mediated induced systemic

resistance. Plant growth-promoting rhizobacteria—Present status and

future prospects. Proceedings of the fourth international workshop on

plant growth-promoting rhizobacteria, Sapporo, Japan. Nakanishi

Printing. pp. 281-282.

Raupach, G.S., Liu, L., Murphy, J.F., Tuzun, S., Kloepper, J.W. 1996. Induced

systemic resistance in cucumber and tomato against cucumber mosaic

cucumovirus using plant growth-promoting rhizobacteria (PGPR). Plant

Dis., 80, 891-894.

Rauthan, B.S., Schnitzer, M. 1981. Effects of a soil fulvic acid on the growth and

nutrient content of cucumber (Cucumis sativus) plants. Plant Soil, 63,

491-495.

Reeve, J.R., Carpenter-Boggs, L., Reganold, J.P., York, A.L., Brinton, W.F.

2010. Influence of biodynamic preparations on compost development and

resultant compost extracts on wheat seedling growth. Bioresour. Technol.,

101(Article in press), 5658–5666.

285
References

Rincon, A., Ruız-Dıez, B., Fernandez-Pascual, M., Probanza, A., Pozuelo, J.M.,

de Felipe, M.R. 2006. Afforestation of degraded soils with Pinus

halepensis Mill.: Effects of inoculation with selected microorganisms and

soil amendment on plant growth, rhizospheric microbial activity and

ectomycorrhizal formation. Appl.Soil Ecol., 34, 42–51.

Robertson, F. 2003. Sugarcane trash management: consequences for soil carbon

and nitrogen. CRC for Sustainable Sugar Production.

Robertson, F.A., Myers, R.J.K., Saffigna, P.G. 1994. Dynamics of carbon and

nitrogen in a long term cropping system and permanent pasture system.

Aust. J. Soil Res., 45, 211-221.

Robertson, F.A., Thorburn, P.J. 2007. Decomposition of sugarcane harvest

residue in different climatic zones. Aust. J. Soil Research, 45, 1-11.

Rodríguez, H., Fraga, R. 1999. Phosphate solubilizing bacteria and their role in

plant growth promotion. Biotechnol Adv., 17(4-5), 319-39.

Romantschuk, M., Sarand, I., Peta¨nen, T., Peltola, R., Jonsson-Vihanne, M.,

Koivula, T., Yrja la, K., Haahtela, K. 2000. Means to improve the effect

of in situ bioremediation of contaminated soil: an overview of novel

approaches. Environ. Pollution, 107, 179–185.

Rousk, J., Brookes, P.C., Bååth, E. 2010. The microbial PLFA composition as

affected by pH in an arable soil. Soil Biol. Biochem., 42, 516-520.

Roy, A.K., Sharma, A., Talukder, G. 1988. Some aspects of aluminium toxicity

in plants. Bot. Rev., 54, 145-177.

Ryan, M. 2003. Compost tea production, application, and benefits, The Rodale

Institute.

286
References

Ryan, M., Wilson, D., Hepperly, P., Travis, J., Halbrendy, N., Wise, A. 2005.

Compost tea potential is still brewing. BioCycle, 46, 30–32.

Ryan, P.R., Ditomaso, J.M., Kochian, L.V. 1993. Aluminium toxicity in roots: an

investigation of spatial sensitivity and the role of the root cap. J. Exp.

Bot., 44, 437-446.

Rynk, R., van de Kamp, M., Willson, G.B., Singley, M.E., Richard, T.L., Kolega,

J.J., Gouin, F.R., Jr, L.L., Kay, D., Murphy, D.W., Hoitink, H.A.J.,

Brinton, W.F. 1992. On-Farm Composting Handbook, Northeast Regional

Agricultural Engineering Service. NY.

Sackenheim, R., Weltzien, H.C., Kast, W.K. 1994. Effects of microflora

composition in the phyllosphere on biological regulation of grapevine

fungal diseases. Vitis, 33, 235–240.

Sample, I. 2007. Global food crisis looms as climate change and population

growth strip fertile land. in: The Guardian.

Scheuerell, S. 2004. Compost tea production practices, microbial properties, and

plant disease suppression. . in: I International Conference on Soil and

Compost Eco-Biology. Spain, pp. 41-51.

Scheuerell, S.J. 2002. Compost teas and compost-amended container media for

plant disease control, Vol. PhD, Oregon State University.

Scheuerell, S.J. 2003. Understanding how compost tea can control disease.

BioCycle, 44(2), 20-25.

Scheuerell, S.J., Mahaffee, W.F. 2000. Assessing aerated and non-aerated watery

fermented compost and Trichoderma harzianum T-22 for control of

powdery mildew (Sphaerotheca pannosa var. rosae) of rose in the

Willamette Valley. Phytopathol., 90, 69.

287
References

Scheuerell, S.J., Mahaffee, W.F. 2004. Compost tea as a container medium

drench for suppressing seedling damping-off caused by Pythium ultimum.

Phytopathol., 94, 1156-1163.

Scheuerell, S.J., Mahaffee, W.F. 2002. Compost tea: principles and prospects for

plant disease control. Compost Sci. Util., 10, 313-338.

Scheuerell, S.J., Mahaffee, W.F. 2006. Variability associated with suppression of

gray mold (Botrytis cinerea) on geranium by foliar applications of non-

aerated and aerated compost teas. Plant Dis., 90, 1201-1208.

Schnitzer, M., Poapst, P. 1967. Effects of a soil humic compound on root

initiation. Nature, 213, 598-599.

Schnurer, J., Rosswall, T. 1982. Fluorescein diacetate hydrolysis as a measure of

total microbial activity in soil and litter. Appl. Environ. Microbiol., 43(6),

1256-1261.

Sebti, K.E.H. 2005. Compost tea effects on soil fertility and plant growth of

organic tomato (Solanum lycopersicum Mill) in comparison with different

organic fertilizers. in: International centre for advanced mediterranean

agronomic studies, Vol. Master Thesis, Istituto Agronomico Mediterraneo

di Bari. Morocco, pp. 67.

Segarra, G., Reis, M., Casanova, E., Trillas, M.I. 2009. Control of powdery

mildew (Erysiphe polygoni) in tomato by foliar applications of compost

tea. J. Plant Pathol., 91(3), 683-689.

Sen, B., Chandra, T.S. 2009. Do earthworms affect dynamics of functional

response and genetic structure of microbial community in a lab-scale

composting system? Bioresour. Technol., 100, 804–811.

288
References

Shannon, E.E., Weaver, W. 1963. The Mathematical Theory of Communication.

The University of Illinois Press, Urbana, Illinois.

Shen, J., Bartha, R. 1996. Priming effect of substrate addition in soil based

biodegradation tests. Appl. Environ. Microbiol., 62, 1428-1430.

Shi, W., Norton, J.M., Miller, B.E., Pace, M.G. 1999. Effects of aeration and

moisture during windrow composting on the nitrogen fertilizer values of

dairy waste composts. Appl. Soil Ecol., 11, 17-28.

Shrestha, K., Shrestha, P., Adetutu, E.M., Walsh, K.B., Harrower, K.M., Ball,

A.S., Midmore, D.J. 2011a. Changes in microbial and nutrient

composition associated with rumen content compost incubation.

Bioresour. Technol., 102, 3848–3854.

Shrestha, K., Shrestha, P., Walsh, K.B., Harrower, K.M., Midmore, D.J. 2011b.

Microbial enhancement of compost extracts based on cattle rumen content

compost - Characterisation of a system. Bioresour. Technol., 102, 8027-

8034.

Siddiqui, Y., Meon, S., Ismail, M.R., Ali, A. 2008a. Trichoderma-fortified

compost extracts for the control of Choanephora wet rot in okra

production. Crop Protec., 27, 385–390.

Siddiqui, Y., Meon, S., Ismail, R., Rahmani, M. 2009. Bio-potential of compost

tea from agro-waste to suppress Choanephora cucurbitarum L. the causal

pathogen of wet rot of okra. Biol. Cont., 49, 38–44.

Siddiqui, Y., Meon, S., Ismail, R., Rahmani, M., Ali, A. 2008b. Bio-efficiency of

compost extracts on the wet rot incidence, morphological and

physiological growth of okra (Abelmoschus esculentus L. Moench).

Scientia Hort., 117, 9-14.

289
References

Singh, A., Sharma, S. 2002. Composting of a crop residue through treatment with

microorganisms and subsequent vermicomposting. Bioresour. Technol.,

85, 107–111.

Singh, P., Suman, A., Tiwari, P., Arya, N., Gaur, A., Shrivastava, A.K. 2008.

Biological pretreatment of sugarcane trash for its conversion to

fermentable sugars. World J. Microbiol. Biotechnol., 24, 667–673.

Skladany, G.J., Metting, F.B. 1992. Bioremediation of contaminated soil. in: Soil

Microbial Ecology: Applications in Agricultural and Environmental

Management, (Ed.) F.B. Metting Jr., Marcel Dekker. New York, pp. 483–

513.

Small, F.G. 2000. Quantifying the socio-economic impacts of harvesting residue

retention systems – Growers‘ survey on burnt and green cane trash

blanket farming systems in the Burdekin and Proserpine districts. Sugar

Research and Development Corporation. BSS173.

Smalla, K., Wachtendorf, U., Heuer, H., Liu, W.T., Forney, L. 1998. Analysis of

Biolog GN substrate utilisation patterns by microbial communities. Appl.

Environ. Microbiol., 64, 1220-1225.

Solomon, D., Lehmann, J., Zech, W. 2000. Land use effects on soil organic

matter properties of chromic luvisols in semi-arid northern Tanzania:

carbon, nitrogen, lignin and carbohydrates. Agric. Ecosyst. Environ., 78,

203–213.

Spaccini, R., Zena, A., Igwe, C.A., Mbagwu, J.S.C., Piccolo, A. 2001.

Carbohydrates in water-stable aggregates and particle size fractions of

forested and cultivated soils in two contrasting tropical ecosystems.

Biogeochem., 53(1), 1-22.

290
References

Sparling, G.P. 1997. Soil microbial biomass, activity and nutrient cycling as

indicators of soil health. in: Biological Indicators of Soil Health, (Eds.)

C.E. Pankhurst, B.M. Doube, V.V.S.R. Gupta, CAB International. New

York, pp. 97–119.

Sparling, G.P., West, A.W. 1988. A direct extraction method to estimate soil

microbial C: calibration in situ using microbial respiration and 14C

labelled cells. Soil Biol. Biochem., 20, 337-343.

Stacey, G., Upchurch, R.G. 1984. Rhizobium inoculation of legumes. Trends in

Biotechnol., 2(3), 65-70.

Stark, C., Condron, L.M., Stewart, A., Di, H.J., Callaghan, M.O. 2007. Influence

of organic and mineral amendments on microbial soil properties and

processes. Appl. Soil Ecol., 35, 79–93.

Steenwerth, K.L., Jackson, L.E., Caldero´n, F.J., Stromberg, M.R., Scow, K.M.

2003. Soil microbial community composition and land use history in

cultivated and grassland ecosystems in coastal California. Soil Biol.

Biochem., 35, 489–500.

Stewart, R.L., Wood, A.W. 1987. Proceedings of green cane symposium. CSR

Ltd.

Suarez-Estrella, F., Vargas-Garcia, M.C., Elorrieta, M.A., Lopez, M.J., Moreno,

J. 2003. Temperature effect on Fusarium oxysporum f. sp. melonis

survival during horticultural waste composting. J. Appl. Microbiol., 94,

475–482.

Sundberg, C., Smårs, S., Jönsson, H. 2004. Low pH as an inhibiting factor in the

transition from mesophilic to thermophilic phase in composting.

Bioresour. Technol., 95(2), 145-150.

291
References

Supanjani, Shim, H.H., Sung, J.J., Lee, K.D. 2006. Rock phosphate-potassium

and rock-solubilising bacteria as alternative, sustainable fertilisers. Agron.

Sustain. Dev., 26, 233-240.

Suresh, A., Pallavi, P., Srinivas, P., Praveen Kumar, V., Chandra, S.J., Reddy,

S.R. 2010. Plant growth promoting activities of fluorescent

pseudomonads associated with some crop plants. Afric. J. Microbiol. Res.,

4(14), 1491-1494.

Suthipradit, S., Edwards, D.G., Ascher, C.J. 1990. Effects of aluminium on tap-

root elongation of soybean (Glycine max), cowpea (Vigna unguiculata)

and green gram (Vigna radiata) grown in the presence of organic acids.

Plant Soil, 124, 233-237.

Sutton, M.R., Wood, A.W., Saffigna, P.G. 1996. Long Term Effects of Green

Cane Trash Retention on Herbert River Soils, (Eds.) J.R. Wilson, D.M.

Hograth, J.A. Campbell, A.L. Garside, CSIRO Division of Tropical Crops

and Pastures. Brisbane, pp. 178-180.

Swift, M.J., Anderson, J.M. 1994. Biodiversity and ecosystem function in

agricultural ecosystems. in: Biodiversity and Ecosystem Function, (Eds.)

E.-D. Schulze, H.A. Mooney. New York, pp. 15–41.

Swift, M.J., van der Meer, J., Ramakrishnan, P.S., Anderson, J.M., Ong, C.K.,

Hawkins, B.A. 1996. Biodiversity and agroecosystem function. in:

Functional Roles of Biodiversity: A Global Perspective, (Eds.) H.A.

Mooney, H. C.J., E. Medina, O.E. Sala, E.D. Schulze, John Wiley and

Sons. Chichester, pp. 261–298.

Sylvia, E.W. 2004. The effect of compost extract on the yield of strawberries and

severity of Botrytis cinerea. J. Sustainable Agri., 25(1).

292
References

Tajuddin, R.M., Ismail, A.F., Salim, M.R. 2004. Effects of oxygen concentration

on microbial growth in aerated palm oil mill effluent using oxygen

enriched air membrane system. In: Regional symposium on membrane

science and technology, Johor Bahru, Johor, Malaysia.

Tan, K.H., Binger, A. 1986. Effect of humic acid on aluminum toxicity in corn

plants. Soil Sci., 141, 20-25.

Tang, J.C., Shibata, A., Zhou, Q., Katayama, A. 2007. Effect of temperature on

reaction rate and microbial community in composting of cattle manure

with rice straw. J. Biosci. Bioeng., 104(4), 321–328.

Taylor, G.J. 1988. The physiology of aluminium tolerance in higher plants.

Commum. Soil Sci. Plant Anal., 19, 1179-1194.

Tejada, M., Gomez, I., Hernandez, T., Garcıa, C. 2010. Response of Eisenia

fetida to the application of different organic wastes in an aluminium-

contaminated soil. Ecotoxicol. Environ. Safety, 73, 1944–1949.

Tejada, M., Gonzalez, J.L., Hernandez, M.T., Garcia, C. 2008. Agricultural use

of leachates obtained from two different vermicomposting processes.

Bioresour. Technol., 99, 6228-6232.

Thompson, J.P. 1996. Correction of dual phosphorus and zinc deficiencies of

linseed (Linumusita tissimum L.) with cultures of vesicular arbuscular

mycorrhizal fungi. Soil Biol. Biochem., 28(7), 941-951.

Thompson, J.P. 1987. Decline of vesicular-arbuscular mycorrhizae in long fallow

disorder of field crops and its expression in phosphorus deficiency of

sunflower. Aust. J. Agric. Res., 38, 847-67.

Thornton, C.M., Cowie, B.A., Freebairn, D.M., Playford, C.L. 2007. The

Brigalow Catchment Study: II. Clearing brigalow (Acacia harpophylla)

293
References

for cropping or pasture increases runoff. Australian Journal of Soil

Research, 45, 496–511.

Tiquia, S.M. 1996. Further composting of pig-manure disposed from the pig-on-

litter (POL) system in Hong, Vol. PhD thesis, The University of Hong

Kong. Pokfulam Road, Hong Kong.

Tiquia, S.M., Tam, N.F.Y., Hodgkiss, I.J. 1996. Microbial activities during

composting of spent pig-manure sawdust litter at different moisture

contents. Bioresour. Technol., 55, 201-206.

Touart, A.P. 2000. Time for compost tea in the Northwest. BioCycle, 41(10), 74-

77.

Tranker, A. 1992. Use of agricultural and municipal organic wastes to develop

suppressiveness to plant pathogens. in: Biological Control of Plant

Diseases, (Eds.) E.S. Tjamos, G.C. Papavizas, R.J. Cook, Plenum press.

NY pp. 35-42.

Tu, C., Ristaino, J.B., Hu, S. 2006. Soil microbial biomass and activity in organic

tomato farming systems: Effects of organic inputs and straw mulching.

Soil Biol. & Biochem., 38, 247–255.

Tuomela, M., Vikman, M., Hatakka, A., Itavaara, M. 2000. Biodegradation of

lignin in a compost environment: a review. Bioresour. Technol., 72, 169 -

183.

Ullrich, S., Karrasch, B., Hoppe, H.-G., Jeskulke, K., Mehrens, M. 1996. Toxic

effects on bacterial metabolism of the redox dye 5-cyano-2,3-ditolyl

tetrazolium chloride. Appl. Environ. Microbial., 62, 4587-4593.

294
References

Unkovich, M., Herridge, D., Peoples, M., Cadisch, G., Boddey, R., Giller, K.,

Alves, B., Chalk, P. 2008. Measuring plant-associated nitrogen fixation in

agricultural systems.

Urban, J., Trankner, A. 1993. Control of gray mold (Botrytis cinerea) with

fermented compost/water extracts. Biological control of foliar and post-

harvest diseases: proceedings of a workshop. West Palearctic Regional

Section. pp. 8-11.

Utkhede, R., Koch, C. 2004. Biological treatments to control bacterial canker of

greenhouse tomatoes. Biocontrol Sci. Technol., 49, 305-313.

Uvarov, A.V., Tiunov, A.V., Scheu, S. 2006. Long-term effects of seasonal and

diurnal temperature fluctuations on carbon dioxide efflux from a forest

soil. Soil Biol. Biochem., 38, 3387–3397.

Vadakattu, G., Paterson, J. 2006. Free living bacteria lift soil nitrogen supply. in:

Farmind Ahead, Vol. 169, pp. 40.

van Elsas, J.D., Garbeva, P., Salles, J.F. 2002. Effect of agronomical measures on

the microbial diversity of soil as related to the suppression of soil-borne

plant pathogens. Biodegradation, 13, 29–40.

van Peer, R., Numann, G.J., Schippers, B. 1991. Induced resistance and

phytoalexin accumulation in biological control of Fusarium wilt of

carnation by Pseudomonas sp. strain WCS417r. Phytopathol., 81, 728-

734.

Vance, E.D., Brookes, P.C., Jenkinson, D.S. 1987. An extraction method for

measuring soil microbial biomass C. Soil Biol. Biochem., 19, 703–707.

295
References

Vásquez, P., Holguín, G., Puente, M.E., López-Cortéz, A., Bashan., Y. 2000.

Phosphate solubilizing microorganisms associated with the rhizosphere of

mangroves in a semiarid coastal lagoon. Biol. Fert. Soils, 30, 460–468.

Vaughan, D., Linehan, D.J. 1976. The growth of wheat plants in humic acid

solutions under axenic conditions. Plant Soil, 44, 445-449.

Vivas, A., Moreno, B., Garcia-Rodriguez, S., Benitez, E. 2009. Assessing the

impact of composting and vermicomposting on bacterial community size

and structure, and microbial functional diversity of an olive-mill waste.

Bioresour. Technol., 100, 1319-1326.

Wakelin, S.A., Colloff, M.J., Harvey, P.R., Marschner, P., Gregg, A.L., Rogers,

S.L. 2007. The effects of stubble retention and nitrogen application on soil

microbial community structure and functional gene abundance under

irrigated maize. FEMS Microbiol. Ecol., 59, 661–670.

Walkley, A., Black, I.A. 1934. An examination of the Degtjareff method for

determining organic carbon in soils: effect of variations in digestion

conditions and of inorganic soil constitutions. Soil Sci., 63, 251-263.

Wardle, D.A., Yeates, G.W., Nicholson, K.S., Bonner, K.I., Watson, R.N. 1999.

Response of soil microbial biomass dynamics, activity and plant litter

decomposition to agricultural intensification over a seven-year period.

Soil Biol. Biochem., 31, 1707–1720.

Watanabe, K. 2001. Microorganisms relevant to bioremediation. Current Opinion

in Biotechnol., 12, 237–241.

Wei, G., Kloepper, J.W., Tuzun, S. 1991. Induction of systemic resistance of

cucumber to Colletotrichum orbiculare by select strains of plant growth-

promoting rhizobacteria. Phytopathol., 81, 1508-1512.

296
References

Welke, S. 2000. Effectiveness of compost extracts as disease suppressants in

fresh market crops in British Columbia. Organic Farming Research

Foundation.

Weltzein, H.C. 1992. Biocontrol of foliar fungal diseases with compost extracts.

in: Microbial Ecology of Leaves, (Eds.) J.H. Andrews, S.S. Hirano,

Springer-Verlag. New York, pp. 430-450.

Weltzein, H.C. 1989. Some effects of composted organic materials on plant

health. Agric. Ecosystems Environ., 27, 439–446.

Weltzein, H.C., Ketterer, N. 1986. Control of Phytophthora infestans on tomato

leaves and potato tubers through water extracts of composted organic

wastes. Phytopathol., 76, 1104.

Weltzien, H.C. 1991. Biocontrol of fungal foliar disease with compost extracts.

Microbial Ecol. Leaves, 430-450.

Weltzien, H.C. 1990. The use of composted materials for leaf disease suppression

in field crops. . in: Monograph – British Crop Protection Council, Vol.

45, pp. 115-120.

Widmer, F., Flieβbach, A., Laczko, E., Schulze-Aurich, J., Zeyer, J. 2001.

Assessing soil biological characteristics: a comparison of bulk soil

community DNA-, PLFA-, and Biolog- analyses. Soil Biol. Biochem., 33,

1029-1036.

Widmer, F., Rasche, F., Hartmann, M., Fliessbach, A. 2006. Community

structures and substrate utilization of bacteria in soils from organic and

conventional farming systems of the DOK long-term field experiment.

Appl. Soil Ecol., 33, 294–307.

Winter, C.K., Davis, S.F. 2006. Organic foods. J. Food Sci., 71(9), 117–124.

297
References

Winterscheidt, H., Minassian, V., Weltzein, H.C. 1990. Untersuchungen zur

biologischen Bekampfung des Falschen Mehltaus der Gurke -

Pseudoperonospora cubensis (Berk. et Curt.) Rost.- mit

Kompostextrakten. Gesunde Pflanzen, 42, 235-238.

Witkamp, M. 1966. Decomposition of leaf litter in relation to environment,

microflora, and microbial respiration. Ecology, 47(2), 194-201.

Wood, A.W. 1991. Management of crop residues following green harvesting of

sugarcane in north Queensland. Soil Till. Res, 20(1), 69-85

Woods, L.E. 1983. Land Degradation in Australia. Department of Home Affairs

and Environment, Canberra: Australian Government Printers.

Wright, R.L. 1989. Soil aluminium toxicity and plant growth. Commun. Soil Sci.

Plant Anal., 20, 1479-97.

Wu, L., Ma, L.Q. 2002. Relationship between compost stability and extractable

organic carbon. J. Environ. Qual., 31, 1323–1328.

Wu, L., Ma, L.W., Martinez, G.A. 2000. Comparison of methods for evaluating

stability and maturity of biosolids compost. J. Environ. Qual., 29, 424–

429.

XLSTAT. 2010. XLSTAT-Pro (Win), Addinsoft, Thierry Fahmy, 40 rue

Damrémont, 75018. Paris, France.

Xu, H. 2000. Effects of a microbial inoculant and organic fertilizers on the

growth, photosynthesis and yield of sweet corn. J. Crop Prod., 3, 183–

214.

Yamada, E., Ozaki, T., Kimura, M. 1998. Determination and behaviour of humic

substances as precursors of trihalomethane in environmental water.

Analytical Sci., 14, 327-332.

298
References

Yan, F., Schubert, S., Mengel, K. 1996. Soil pH increase due to biological

decarboxylation of organic anions. Soil Biol. Biochem., 28(4), 611-624.

Yohalem, D.S., Harris, R.F., Andrews, J.H. 1994. Acquous extracts of spent

mushroom substrate for foliar disease control. Compost Sci. Util., 2, 67-

74.

Young, I.M., Ritz, K. 2000. Tillage, habitat space and function of soil microbes.

Soil Tillage Res., 53, 201-213.

Zak, D.R., Holmes, W.E., MacDonald, N.W., Pregitzer, K.S. 1999. Soil

temperature, matric potential, and the kinetics of microbial respiration and

nitrogen mineralization. Soil Sci. Soc. Am. J., 63, 575-584.

Zak, J.C., Willig, M.R., Moorhead, D.L., Wildman, H.G. 1994. Functional

diversity of microbial communities: a quantitative approach. Soil Biol.

Biochem., 26, 1101–1108.

Zhang, W., Han, D.Y., Dick, W.A., Davis, K.R., Hoitink, H.A.J. 1998. Compost

and compost water extract-induced systemic acquired resistance in

cucumber and arabidopsis. Phytopathol., 88(5), 450-455.

Zmora-Nahum, S., Danon, M., Hadar, Y., Chen, Y. 2008. Chemical properties of

compost extracts inhibitory to germination of Sclerotium rolfsii. Soil Biol.

Biochem., 40, 2523–2529.

Zwieten, M.V., Stovold, G., Zwieten, L.V. 2007. Alternatives to copper for

disease control in the Australian organic industry. RIRDC publication.

299

You might also like