estudio de la dinámica microbiana durante la fase de

194
Iria Villar Comesaña 2017 TESE DE DOUTORAMENTO Estudio de la dinámica microbiana durante la fase de maduración del compostaje de residuos orgánicos. Vermicompostaje como alternativa de tratamiento

Upload: others

Post on 15-Oct-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Iria Villar Comesaña

2017

Iria

Villa

r C

omes

aña .

TE

SE

DE

DO

UT

OR

AM

EN

TO

TESE DE DOUTORAMENTO

Estudio de la dinámica microbiana durante

la fase de maduración del compostaje de

residuos orgánicos.

Vermicompostaje como alternativa de

tratamiento

Est

udi

o de

la d

inám

ica m

icro

biana d

ura

nte

la f

ase

de

madu

raci

ón d

el c

ompos

taje

de

resi

duos

org

ánic

os.

Ver

mic

ompos

taje

com

o alt

ernati

va d

e tr

ata

mie

nto

2017

Escola Internacional de Doutoramento

Iria Villar Comesaña

TESE DE DOUTORAMENTO

Estudio de la dinámica microbiana durante la fase de maduración

del compostaje de residuos orgánicos. Vermicompostaje como

alternativa de tratamiento

Dirixida polo doutor:

Salustiano Mato de la Iglesia

2017

O meu neno e a miña avoa

AGRADECIMIENTOS

Mi agradecimiento más sincero a Salustiano Mato por compaginar su exigente trabajo de

rector con la dirección de esta tesis dándome la oportunidad de conocer el mundo de la

investigación, ayudándome y destinando su tiempo libre a diseñar, discutir y enriquecer los

estudios de esta tesis.

A Josefina Garrido por escuchar y recibirme en su despacho siempre con una sonrisa y un

buen consejo.

A todos los compañeros del laboratorio, que han sido muchos, y a los que les debo un poco de

lo aprendido en todos estos años: Pachila, Jorge, Velando, Juan, Cereijo, Emma, Elisa, Elena,

Rubén, Salomé, Andrea, Manuel, Cristóbal, Alberto L., María, Cristina, Pablo, Jessica, Lorena,

Hugo, César, Amaia, Sarr, Álex, Yeon, Álvaro, Alberto DS y Javi. Gracias a todos.

A Emilio por echar una mano siempre que lo he necesitado y, de manera muy especial, a

Domingo por sus clases magistrales de informática y automática y todo su apoyo, consejo y

ayuda. Gracias a él sé diferenciar una sonda termopar de una pt100.

Gracias al personal de los servicios del CACTI, del departamento de Ecología y Biología

Animal, de Leboriz SA. y de FEUGA.

A todas las empresas, instituciones y personas con las que he trabajado todos estos años en

este mundo de la gestión de residuos: Vigozoo, ROS ROCA S.A., Complejo Medioambiental

Sierra del Barbanza, Codisoil S.A., Applus Agroambiental S.A., Compost Galicia S.A., Leal &

Soares S.A., Ayuntamiento de Allariz, Cofradía de Pescadores de Vilaboa, Ecocelta, Ramón

Plana, CETMAR, REGATA y Eurofins Agroambiental S.A.

Gracias a Jose de la depuradora de Cangas do Morrazo, a la empresa FRINOVA S.A., a la Granja

El Pego y Compost Galicia S.A. por suministrar los residuos y materiales para la presente

tesis.

Mil gracias a David por acompañarme y animarme en este periplo, por tus críticas, tus

consejos y tus elogios. Sin ti nada de esto hubiera sido posible. Siempre a mi lado.

Al pequeño Edu, mi mayor alegría. Todo es posible a tu lado mi niño.

A mi queridísima abuela, a mis padres Carlos y Rosa y mis hermanas Ana y Olalla, por ser mi

ejemplo a seguir y por vuestra inmensa paciencia y apoyo. A Roberto, Rubén, Iker y Carlota. A

Suso, Fini y Jacobo por recibirme en vuestra casa. Gracias familia.

A mis amigas Noe y Patri por haber estado siempre a mi lado a pesar de no poder vernos

tanto. A Lorena SC por nuestras charlas de café.

A todos y todas los que habéis colaborado con un granito en esta pila de compost.

Gracias

ÍNDICE

Introducción………………………………………………………………………………………………………………………………. 1

1. Degradación de compuestos orgánicos………………………………………………………………………... 3

2. Compostaje…………………………………………………………………………………………………………………. 6

2.1 Generalidades…………………………………………………………………………………………………… 6

2.2 Fases del proceso de compostaje………………………………………………………………………... 7

2.3 Condiciones previas y control del proceso………………………………………………………….. 8

2.4 Microorganismos implicados en el compostaje………………………………............................ 17

2.5 Sistemas de compostaje……………………………………………………………………………………... 20

3. Vermicompostaje………………………………………………………………………………………………………… 25

2.1 Generalidades…………………………………………………………………………………………………… 25

2.2 Organismos implicados en el vermicompostaje………………………………………………….. 25

2.3 Condiciones previas y control del proceso………………………………………………………….. 30

2.4 Sistemas de vermicompostaje……………………………….............................................................. 41

Objetivos…………………………………………………………………………………………………………………………………..... 47

Capítulo 1…………………………………………………………………………………………………………………………………… 51

“Evolution of microbial dynamics during the maturation phase of the composting of different

types of waste”

Capítulo 2…………………………………………………………………………………………………………………………………… 65

“Seafood-processing sludge composting: changes to microbial communities and physico-chemical

parameters of static treatment versus for turning during the maturation stage”

Capítulo 3…………………………………………………………………………………………………………………………………… 85

“Changes in microbial dynamics during vermicomposting of fresh and composted sewage sludge”

Capítulo 4…………………………………………………………………………………………………………………………………… 113

“Product quality and microbial dynamics during vermicomposting and maturation of compost

from pig manure”

Discusión general……………………………………………………………………………………………………………………... 145

Conclusiones………………………………………………………………………………………………………………………………. 157

Bibliografía………………………………………………………………………………………………….......................................... 161

INTRODUCCIÓN

Introducción

3

1. DEGRADACIÓN DE COMPUESTOS ORGÁNICOS

Como consecuencia de las actividades habituales del ser humano se generan una gran

cantidad y volumen de residuos de diferentes orígenes y que varían en composición y

peligrosidad. En 2014, la cantidad total de residuos generados en la EU-28 por la totalidad de

actividades económicas y hogares ascendió a 2598 millones de toneladas. La cantidad media

generada por habitante en España, en ese mismo año, ascendió a 2387kg habitante-1

descendiendo el ratio a 1435kg habitante-1 excluyendo los principales residuos minerales

(Eurostat, 2017). Una parte importante de estos residuos son materiales biodegradables que

presentan diferentes características y composición dependiendo, fundamentalmente, de su

origen biológico: agrícola, ganadero, forestal, agroindustrial o urbano. En general, estos

residuos no deben ser vertidos al suelo de manera directa debido a que su descomposición no

controlada y su toxicidad pueden generar efectos perjudiciales. Estos residuos

biodegradables poseen materia orgánica no estabilizada, por lo que presentan un elevado

nivel de fitotoxicidad y, frecuentemente, una alta carga de agentes patógenos, como virus,

bacterias, hongos y parásitos, tanto para los seres humanos como para los animales y plantas,

y siendo, por lo tanto, un riesgo para el medioambiente y para la salud. De esta manera, los

residuos orgánicos deben ser estabilizados y acondicionados reduciendo la emisión de olores

y los posibles riesgos anteriormente citados antes de su disposición al suelo.

La materia orgánica presente en el suelo de manera natural está formada,

principalmente, por la biomasa viva y por los restos vegetales y animales muertos que se

depositan en el suelo. Esta biomasa está sometida a una sucesión de transformaciones físicas,

químicas y biológicas. La descomposición en el suelo de los compuestos orgánicos se

desarrolla a través de la degradación gradual de componentes como glúcidos, polisacáridos,

proteínas, lípidos, otros materiales alifáticos y ligninas, llevada a cabo por la acción

enzimática de la macrofauna, mesofauna y microfauna que habitan en el suelo en una

interacción constante. Esta degradación de la materia orgánica conlleva un proceso de

mineralización hasta obtener nutrientes que puedan ser asimilados por las plantas. A su vez,

la materia orgánica no mineralizada sufre un proceso de transformación que, junto con

moléculas orgánicas sintetizadas por el metabolismo anabólico, forman los compuestos

húmicos, es decir, polímeros orgánicos de difícil degradación. Además, de estos procesos de

mineralización y humificación, se sucede también la síntesis de secreciones diversas como

vitaminas, fitoreguladores, antibióticos, etc. (Senesi, 1989). La velocidad del proceso de

mineralización de la materia orgánica en el suelo está condicionada por factores internos

propios de su naturaleza y otros externos como la temperatura, la humedad, la luz, etc.

Introducción

4

(Curtin et al., 1998; Kirschbaum, 1995; Leirós et al., 1999). Los descomponedores primarios

del suelo son los principales responsables de la rotura física, dispersando y mezclando la

materia orgánica por el suelo, alterando enzimáticamente los componentes, excretando heces

y fluidos, ingiriendo la propia biomasa microbiana del suelo y, como consecuencia,

exponiendo las moléculas orgánicas a los descomponedores secundarios. Dentro de la

mesofauna y macrofauna del suelo se encuentran especies de miriápodos, arácnidos, insectos,

crustáceos, anélidos, etc, que se alimentan directamente de la materia orgánica y de otra

fauna vinculada al suelo. La descomposición secundaria es desarrollada por bacterias y

hongos que producen las enzimas responsables de la degradación de la materia orgánica. Los

microorganismos del suelo sintetizan y liberan enzimas extracelulares para hidrolizar

sustratos de peso molecular elevado hasta que sean lo suficientemente pequeños para ser

metabolizados por sus células (Coleman et al., 2004; Lavelle and Spain, 2001; Senesi et al.,

2009). Así, durante las primeras fases de la descomposición de los materiales orgánicos,

aparece la microbiota asociada a la degradación de los compuestos solubles de fácil

asimilación y de bajo peso molecular. Por lo general, se trata de microorganismos estrategas

de la r con ventajas competitivas debido a sus altas tasas de crecimiento y reproducción que

dan lugar a una importante proliferación microbiana en los sustratos colonizados cuando los

recursos son abundantes. En consecuencia, los sustratos fácilmente asimilables se reducen y

las poblaciones colonizadoras iniciales decrecen, pudiendo desarrollarse formas de

resistencia. La microfauna especializada en degradar compuestos más complejos continúa la

descomposición de la materia orgánica a un ritmo más lento. Los hongos son los más

eficientes en la degradación de compuestos de lignina y celulosa, especialmente, hongos de la

podredumbre blanca y parda (tanto basidiomicetes como ascomicetes). Además, suelen ser

dominantes en condiciones ácidas debido a la falta de competencia microbiana por los

recursos disponibles (Boer et al., 2005; Senesi et al., 2009). La mayoría de los

microorganismos del suelo son aerobios, por lo que, cuando las condiciones de oxígeno son

limitantes, la actividad microbiana se ve afectada y se produce un cambio hacia poblaciones

microbianas con actividad anaerobia o fermentativa que son procesos menos eficientes en la

degradación de la materia orgánica (Coyne, 1999). En el caso de que las condiciones

conlleven una falta de oxígeno, las bacterias son las principales responsables de la

degradación.

Killham (1994) expone que en un suelo de pradera la biomasa aproximada de bacterias

y hongos es de 1-2 y 2-5 t ha-1, respectivamente. Las bacterias que se encuentran con mayor

frecuencia pertenecen a los géneros Bacillus y Pseudomonas, siendo Streptomyces,

Arthrobacter, Nocardia y Actinoplanes algunas de las actinobacterias aisladas más

Introducción

5

importantes. Dentro de los géneros más comunes de hongos aislados se encuentran

Mortariella, Mucor, Chaetomium y Penicillium, por ejemplo. El tipo y abundancia de bacterias

y hongos depende de muchos factores como la disponibilidad de nutrientes y las condiciones

medioambientales en el suelo (Pierzynski et al., 2005). Como anteriormente se ha comentado,

los microorganismos son los principales descomponedores de la materia orgánica pero no los

únicos. La accesibilidad del sustrato para los microorganismos puede ralentizar la tasa de

descomposición, por lo que la acción de la restante fauna del suelo tiene un papel

fundamental. Las lombrices de tierra son, sustancialmente, el grupo más importante de la

fauna presente en el suelo, destacando su papel en términos de formación y mantenimiento

de la estructura del medio, además de mejorar la fertilidad del suelo. Aunque no son

dominantes en número, su tamaño hace que sean uno de los mayores contribuidores en la

biomasa de invertebrados del suelo y su actividad favorece al mantenimiento de la fertilidad

en distintos tipos de suelo (Edwards, 2004).

Puesto que los residuos orgánicos son ricos en nutrientes y materia orgánica, es

razonable que sean devueltos al suelo en condiciones óptimas para garantizar su fertilidad y

completar el ciclo natural. Del mismo modo que sucede en la naturaleza con los ciclos de

nutrientes, los residuos orgánicos pueden sufrir procesos de biodegradación, es decir, la

rotura de los componentes más complejos en compuestos simples mediante la acción de

organismos vivos. Así mismo, la acción de los organismos provoca la síntesis desde moléculas

simples a moléculas más complejas de mayor estabilidad. Dentro de los procesos que pueden

ser empleados para la estabilización de compuestos orgánicos se encuentran el compostaje y

el vermicompostaje que se describirán en los siguientes apartados de la presente tesis.

Introducción

6

2. COMPOSTAJE

2.1. Generalidades

Desde un punto de vista etimológico compostaje deriva del latín compositum que

significa compuesto o mezcla. Así, el compostaje hace referencia a la biodegradación de una

mezcla de sustratos en estado sólido y en condiciones aerobias llevada a cabo por

comunidades microbianas compuestas por diversas poblaciones. Es un proceso espontáneo y

exotérmico que provoca la liberación de energía en forma de calor causando la elevación de la

temperatura hasta condiciones termófilas (Insam and de Bertoldi, 2007). Durante el proceso

de compostaje se generan diferentes condiciones térmicas que llevan a la sucesión de

distintas comunidades microbianas, así mismo, a lo largo de esta biodegradación se produce

la liberación de sustancias tóxicas (fitotoxinas) como los compuestos intermediarios

amoníaco, ácidos orgánicos de bajo peso molecular y el óxido de etileno. La definición que

Zucconi and De Bertoldi (1987) ofrecieron a la comunidad científica es una de las

descripciones más completas de compostaje: “proceso biooxidativo controlado, en el que un

sustrato orgánico heterogéneo en estado sólido experimenta una etapa termofílica y una

liberación transitoria de fitotoxinas, obteniéndose como productos dióxido de carbono, agua,

minerales y materia orgánica estabilizada denominada compost”. Los términos que destacan

en esta definición son: sólido, aerobio, termófilo y controlado. Estos cuatro términos son

básicos y necesarios para que un proceso de degradación biológica de material orgánico sea

considerado como compostaje. Estos autores incluyeron en la definición el término

controlado, haciendo alusión a la necesidad de monitorización de los parámetros tales como

la temperatura, oxígeno, humedad, etc. a lo largo del proceso, así como, de controlar la

composición y naturaleza del sustrato heterogéneo en fase de compostaje. De esta manera, se

diferencia un proceso natural de degradación de sustratos orgánicos o una putrefacción que

se sucede habitualmente en acumulaciones sobre el suelo, pilas de estiércol o vertederos,

frente al proceso de compostaje. En la Figura 1 se pueden observar los distintos productos

que se generan durante el proceso de compostaje.

Introducción

7

Figura 1. Esquema del proceso de compostaje (elaboración propia).

Los cambios fisicoquímicos y biológicos que se suceden durante el compostaje dan

como resultado la generación de calor, la formación de dióxido de carbono, agua y, como

principal producto, el compost formado por tres componentes básicos: materia orgánica

estabilizada, materia mineral y microbiota. Por lo tanto, la principal finalidad del compostaje

es la obtención de un producto final de uso agronómico a partir del tratamiento de residuos

orgánicos y, por lo tanto, es una tecnología de valorización medioambientalmente respetuosa

al transformar desechos en productos aplicables al suelo posibilitando el cierre del ciclo de

los nutrientes y el desarrollo de la economía circular de la región.

2.2. Fases del proceso de compostaje

Uno de los factores principales y determinantes del compostaje es la temperatura que

conduce el desarrollo del proceso. Básicamente, se considera que el compostaje en óptimas

condiciones puede dividirse en 4 fases:

Fase mesófila inicial (desde la temperatura ambiental hasta los 45 °C): en esta etapa

comienza la degradación de los compuestos más fácilmente biodegradables, como azúcares,

grasas, almidón y proteínas, que tiene como consecuencia el aumento de la actividad

microbiana y la elevación de la temperatura. Esta etapa puede durar desde horas hasta varios

días.

Fase termófila (45-70ºC): a medida que se eleva la temperatura los organismos

adaptados a temperaturas altas reemplazan a los organismos mesófilos. De esta manera,

continúa la degradación de los compuestos más lábiles desarrollándose una actividad

metabólica elevada que genera la prolongación de las altas temperaturas. En esta etapa

comienza la degradación de los compuestos más complejos y resistentes por parte de la

Introducción

8

microbiota termófila. Debido a las elevadas temperaturas se produce la higienización del

sustrato al eliminarse los patógenos, parásitos y semillas de malas hierbas. Esta etapa puede

durar desde varios días a incluso meses dependiendo del tipo de residuo, las características

del sistema y el control sobre el proceso.

Fase de enfriamiento: el agotamiento de los materiales fácilmente biodegradables

conlleva la disminución de la actividad microbiana y, con ello, el descenso térmico hasta

temperaturas ambientales, produciéndose la recolonización por parte de la microbiota

mesófila con capacidad para degradar compuestos como la celulosa, hemicelulosa y otros

polímeros. Esta fase puede durar desde varios días a semanas por eso, en ocasiones, la fase de

enfriamiento y maduración se consideran como una única etapa del proceso.

Fase de maduración: el material orgánico se estabiliza, predominando los procesos de

degradación de los compuestos más recalcitrantes y la polimerización de sustancias similares

al humus. Esta etapa del proceso suele requerir varios meses.

Dependiendo de la naturaleza del material de partida, las condiciones ambientales, el

sistema de compostaje, las operaciones realizadas durante el proceso y la calidad de compost

requerida, la duración del compostaje es variable y puede prolongarse desde los 3-4 meses

hasta el año.

2.3. Condiciones previas y control de proceso

El proceso de compostaje es dirigido por la actividad de los microorganismos, por lo

que los factores que puedan limitar el desarrollo de la microbiota particular en cada fase,

afectarán al desarrollo del proceso de compostaje. Por ello, el compostaje requiere una serie

de requisitos previos y el control de las condiciones que se suceden a lo largo del mismo.

2.3.1. Acondicionamiento

Antes de proceder a compostar un residuo ha de tenerse en cuenta una serie de

parámetros previos que van a condicionar el desarrollo de la actividad microbiana. Las

características del material inicial determinan la evolución del proceso de compostaje, su

eficacia y la calidad del compost. A continuación se presentan algunos de los parámetros más

importantes:

Introducción

9

� Espacio de aire libre y tamaño de partícula

La matriz de compostaje es una masa de partículas sólidas que debe contener

suficientes poros e intersticios para posibilitar que se desarrolle un proceso aerobio, de

manera que, el aire circule por el interior de la masa aportando una concentración óptima de

oxígeno, eliminando el dióxido de carbono y el exceso de humedad y limitando la

acumulación excesiva de calor (Haug, 1993). Por lo tanto, si el residuo no presenta la

estructura óptima, como ocurre, por ejemplo, con residuos pastosos como los lodos de

depuradora urbana o los purines de cerdo, es necesario adicionar un material que aporte

estructura. Por lo general, se emplean como agentes estructurantes materiales de naturaleza

orgánica como los residuos agroforestales. En la Figura 2 se exponen imágenes de materiales

estructurantes.

Figura 2. A la izquierda madera triturada y a la derecha restos de paja que pueden ser

empleados como agente estructurante.

La porosidad o espacios de aire en la matriz de compostaje puede ser medida mediante

diversos métodos. Uno de los más utilizados es la medición del espacio libre de aire (free air

space, FAS). Los valores mínimos de FAS en un proceso de compostaje para asegurar la

actividad biológica se sitúan en torno al 30%, mientras que, los valores del 60-70% parecen

ser excesivos para alcanzar temperaturas termófilas en residuos con bajo contenido de

materia orgánica biodegradable (Haug, 1993; Ruggieri et al., 2009). El uso de agentes

estructurantes posibilita alcanzar valores de FAS óptimos en el sustrato inicial, aunque estos

agentes también afectan a los balances de nutrientes y humedad, como se verá a

continuación.

Se han realizado múltiples investigaciones que han estudiado el ratio residuo/agente

estructurante para determinar su influencia en el proceso de compostaje así como el tipo de

material estructurante y su tamaño de partícula (Tabla 1). El ratio óptimo para el desarrollo

del compostaje va a depender de factores como las condiciones de proceso (sistema de

compostaje, volumen, tiempo, etc.), el residuo (pastosidad, biodegradabilidad, etc.), el origen

Introducción

10

del agente estructurante, el tipo de estructurante (fresco o recirculado), el tamaño de

partícula, las condiciones de la mezcla (FAS, relación C/N, humedad), etc.

Tabla 1. Referencias de investigaciones acerca del material estructurante tipología, ratio e

influencia sobre el proceso de compostaje

Residuo Agente estructurante Sistema de compostaje

Referencia

Estiércol

Paja picada de cebada 1:2 (v:v)

Turba Sphagnum y paja en proporciones de

130: 20: 5 (f.w.)

Tambor

horizontal

giratorio 25m3

(Vuorinen,

2000)

Lodo de

depuradora

urbana

Astillas de madera

Tamaños partícula 20,10,5mm

Proporción 1:1, 2:1, 4:1 (v:v)

Vasos Dewar 45L

Reactor estático

100L

(Gea et al.,

2007)

Lodo de

depuradora

urbana

Triturado Acacia sp 40 mm

5 proporciones (w:w):1/0, 1/1, 1/2, 1/3

Reactor

35kg

(Yañez et al.,

2009)

Lodo de

depuradora

matadero

Residuos verdes y pallets madera triturados.

1/6 (d.w.)

5 tamaños partícula 8, 12, 20, 30, 40 mm

Reactor

cilíndrico 300L

(Trémier et

al., 2009)

Residuos de

alimentos

Cáscara de arroz, serrín y salvado de arroz

Ajuste ratio en función de la humedad, C/N y

espacio aire

Reactor

cilíndrico 180L

(Chang and

Chen, 2010)

Estiércol de

cerdo

Serrín (4:1 w:w), desechos verdes triturados

(4:1 w:w), paja picada (4:1 w:w)

astillas de madera (4:1 w:w)

Pilas volteadas

47.5kg

(Nolan et al.,

2011)

Lodo de

depuradora

urbana

Pallets de madera frescos y recirculados

1/3 (v/v)

Tamaños partícula: >20,<20 mm

2 humedades 50% 65%

Reactores

cilíndricos 47L

(Huet et al.,

2012)

Digestato de

purín de cerdo

Paja de trigo, bagazo, podas de vid, poda de

pimiento; 10 mm

Distintas proporciones (d.w.)

Reactor 350L con

volteos

(Bustamante

et al., 2013)

Residuos sólidos

municipales

Astillas de madera (1:1, v:v)

Tubos de polietileno (1:1, v:v)

Reactor aireación

forzada 50L

(Maulini-

Duran et al.,

2014)

Los microorganismos acceden al residuo y se alimentan de él a una velocidad que va a

depender de la relación superficie/volumen de las partículas de la matriz de compostaje. De

esta manera, partículas de tamaño elevado requerirán más tiempo de ataque microbiano y

ralentizarán el proceso de compostaje mientras que partículas pequeñas reducen la

porosidad, y aumentan la compactación y la posibilidad de anaerobiosis. Según Diaz et al.

(2007) el tamaño de partícula depende de la naturaleza física del residuo que, para

materiales que no se compactan con facilidad, se considera entre 1-5 cm. Por ejemplo,

Introducción

11

Hamoda et al. (1998) mostraron que para el compostaje de residuo sólido municipal el

tamaño de partícula óptimo es de 4 cm.

� Humedad

Debido a que el compostaje es un proceso de degradación biológica, es importante que

el contenido de agua disponible sea suficiente para hacer frente a los requerimientos

fisiológicos de la microbiota. El agua actúa como medio de transporte, no sólo para las

sustancias solubles con las que se alimentan los microorganismos sino también, para eliminar

los productos de desecho resultado del metabolismo celular. El contenido óptimo de

humedad al inicio del proceso depende del tipo de material a compostar, pero varía entre 50–

70% (Bernal et al., 2009; de Bertoldi et al., 1996; Richard et al., 2002). Contenidos mayores de

humedad pueden provocar que el agua ocupe los microporos de la mezcla dificultando la

oxigenación, mientras que, contenidos menores provocan el descenso de la actividad

biológica. En la Tabla 2 se presentan las humedades máximas recomendadas para distintos

tipos de residuos. Así mismo, la adición de materiales estructurantes afecta a la humedad,

generalmente, provocando la absorción de agua del sustrato si se emplean estructurantes

porosos y no saturados en agua (Haug, 1993).

Tabla 2. Contenidos de humedad recomendados para el compostaje. Modificado de Haug,

(1993) y Diaz et al. (2002).

Tipo de residuo Humedad (%)

Paja 70-85

Madera en serrín o pequeñas virutas 80-90

Cáscara de arroz 70-85

Residuo sólido urbano 55-65

Estiércoles 55-65

Lodo digerido o fresco 55-60

Desechos húmedos (recortes de hierba, basura, etc) 50-55

� Relación C/N

Se considera que los microorganismos vivos emplean para su metabolismo 30 partes

de carbono por cada parte de nitrógeno, requiriendo carbono como fuente de energía y como

componente de las células y nitrógeno para la síntesis proteica y los ácidos nucleicos. Se

considera que la matriz de compostaje es adecuada cuando presenta un ratio inicial de C/N

entre 25-35 (Bernal et al., 2009; Diaz and Savage, 2007; Haug, 1993; Poincelot, 1972). Valores

elevados del ratio C/N, como ocurre en residuos agroforestales con altos contenidos de

carbono, implican un descenso de la actividad biológica por falta de nitrógeno para el

Introducción

12

metabolismo y, por lo tanto, una ralentización del proceso. Valores bajos del ratio C/N, como

ocurre en el compostaje de lodos de depuradora de aguas residuales urbanas, sufren un

exceso de nitrógeno que se puede perder por volatilización en forma de amoníaco o por

lixiviación en forma de amonio. La adicción de material estructurante ayuda a mejorar este

ratio al proveer de carbono orgánico a la mezcla. Sin embargo, el ratio C/N óptimo está

condicionado por la naturaleza del residuo, por lo que, si el carbono forma parte de

compuestos de difícil degradación, como la lignina, estará lentamente disponible para los

microorganismos (Diaz and Savage, 2007). Por ello es importante considerar el ratio en

función del C y N más asimilable o biológicamente disponible.

Por lo tanto para obtener una matriz en la que se suceda un proceso de compostaje

optimizado, es necesario acondicionar el residuo tanto física, química como

termodinámicamente (Haug, 1993).

− Físico o estructural: acondicionar el residuo para eliminar limitaciones causadas por

la falta de humedad o porosidad.

− Químico: acondicionar el residuo para eliminar limitaciones causadas por la falta de

nutrientes u otros químicos desbalanceados.

− Termodinámico: acondicionar la mezcla para asegurar la disponibilidad de energía

suficiente para conducir el proceso.

En ocasiones, se utilizan materiales estructurantes recirculados, es decir, ya empleados

en otros ciclos de compostaje y recuperados, que van a actuar como inóculo microbiano,

además de la mejora física y química que ejercen.

En las plantas de compostaje no siempre es posible acondicionar el residuo de la

manera más idónea y, en muchas ocasiones, los niveles de nutrientes, en concreto el ratio

C/N, no se encuentran en los valores considerados óptimos, por lo que el proceso ha de ser

controlado para reducir al máximo las pérdidas de nitrógeno.

2.3.2. Seguimiento del proceso

Una vez establecidos y corregidos los parámetros relativos a la naturaleza del sustrato,

es necesario el seguimiento y control del proceso dentro de valores adecuados para cada fase

del compostaje.

Introducción

13

Temperatura

La evolución de la temperatura a lo largo del tiempo es un parámetro clave a

consecuencia de la elevada actividad microbiana y, por lo tanto, con relación directa en la

degradación de la materia orgánica. Un proceso de compostaje se desarrolla en tres fases

térmicas: mesófila inicial, termófila y regreso a valores mesófilos. En la Figura 3 se pueden

observar dos ejemplos de evoluciones de temperatura en distintos sistemas y materiales. Los

diferentes perfiles observables son fruto de la distinta naturaleza del sustrato, del volumen

de residuo tratado y del sistema de compostaje empleado.

Figura 3. A la izquierda, evolución de la temperatura en una pila volteada de estiércol de 40m3

durante 230 días y a la derecha, evolución de la temperatura en un reactor estático de 600L con

ventilación forzada durante 22 días.

Para una correcta higienización del material y la obtención de un compost libre de

parásitos y semillas de malas hierbas se han expuesto, por ejemplo, que se han de mantener

en continuo temperaturas por encima de 55ºC durante 15 días (European Commission,

2001). Siendo el exceso de temperatura por encima de los 70ºC no deseable ya que puede

provocar la muerte de la mayoría de los microorganismos, retardar la colonización en las

fases posteriores y, como consecuencia, retrasar la degradación del residuo. De esta manera,

la temperatura durante el proceso debe ser controlada asegurando que se alcancen

temperaturas termófilas y que se sostengan en el tiempo lo suficiente para garantizar la

higienización pero sin exceder de los 70ºC. El control de temperatura se puede realizar o

bien, para enfriar la masa o bien, para reactivar el proceso y se suele efectuar mediante

(Bernal et al., 2009):

− Un sistema de control de retroalimentación de temperatura, mediante el cual se

introduzca aire a la masa en compostaje por ventilación controlada.

Introducción

14

− Operaciones de volteo que consisten en la mezcla de la masa en compostaje mediante

medios mecánicos que provocan la dispersión del calor y el enfriamiento pero que

también pueden reactivar el proceso al redistribuir los nutrientes.

− Control del tamaño y la forma de la masa en proceso de compostaje.

Oxígeno

Mantener los niveles de oxígeno es un factor clave para que se desarrolle un proceso

biológico aerobio. La concentración de oxígeno en la masa de compostaje no debe ser inferior

al 5% (Epstein, 2011), ya que se produciría una sucesión hacia microorganismos anaerobios

y, por lo tanto, hacia procesos de fermentación no deseables y la generación de malos olores.

Para el control de los niveles de oxígeno durante el compostaje ha de controlarse la aireación

de la masa. Además de satisfacer la demanda de oxígeno para la descomposición orgánica, la

aireación también favorece la regulación de excesos de agua y ayuda a mantener la

temperatura en valores adecuados (Haug, 1993). La aireación puede ser proporcionada

mediante sistemas de ventilación, y operaciones de volteo o ambos. Así mismo, también se

puede producir la aireación pasiva de manera natural cuando tanto la estructura y porosidad

de la mezcla, así como, la forma y tamaño de la pila posibilitan el intercambio de gases. Esta

aireación natural pasiva se produce cuando el incremento de temperatura provoca el

calentamiento del aire, por lo que la diferencia entre la densidad del aire exterior menos

húmedo con el aire del interior de la pila más húmedo y caliente, genera un ascenso del aire

hacia la parte superior y la entrada de aire frío desde la parte inferior (Figura 4). Los volteos

u otras acciones mecánicas incrementan el aire libre, disminuyen el tamaño de partícula y

minimizan los efectos de la compactación posibilitando condiciones en la pila que mejoran la

ventilación natural.

Figura 4. Esquema de la ventilación natural en una pila de compostaje (Haug, 1993).

Introducción

15

Humedad

El proceso de compostaje tiene como efecto asociado la pérdida de agua como

consecuencia de su evaporación debido a las altas temperaturas, de manera que, los valores

son más altos al inicio y decrecen a medida que el proceso avanza. El requerimiento de

humedad es mayor durante las fases iniciales cuando la actividad es alta, considerando un

valor límite de 30-35% por debajo del cual se inhibe el proceso. Existen una serie de

actuaciones que provocan el descenso de la humedad como son, la aireación forzada de la

masa y los volteos, que pueden utilizarse para descender la humedad en caso de exceso (de

Bertoldi et al., 1996). Así mismo, debe incorporarse agua al proceso en caso de que los

valores de humedad sean limitantes. Es importante destacar que usualmente es ventajoso

obtener un compost con niveles de humedad bajos que faciliten el ensacado y el transporte,

de manera que los valores oscilen entre 35-40%. El Real Decreto de fertilizantes y afines

(Boletín Oficial del Estado, 2013) establece un contenido máximo del 40% para considerar un

compost como enmienda orgánica.

pH

La mayoría de los microrganismos no sobreviven a pH inferiores a 3, ni superiores a 11,

por lo que residuos entre 3-11 pueden considerarse como potencialmente compostables.

Usualmente, el pH no es un factor clave para el compostaje ya que muchos residuos se

encuentran en rangos de pH entre 5.5-8.0 que son estimados como óptimos (de Bertoldi et al.,

1983). Sin embargo, pH bajos provocan la inhibición de los microorganismos termófilos y,

por lo tanto, un retraso en el paso de condiciones mesófilas a termófilas (Sundberg et al.,

2004) y pH básicos, junto con condiciones de elevadas temperaturas, provocan la

volatilización del amonio. En general, se considera que el pH sigue una evolución en fases, por

lo que puede ser considerado un buen indicador de la degradación de la materia orgánica:

condiciones ácidas durante los primeros días, como consecuencia de la degradación de

polisacáridos fácilmente hidrolizados y nueva síntesis de ácidos orgánicos simples; a medida

que la temperatura aumenta, el pH aumenta hasta 8-9, principalmente debido a la

degradación metabólica de los ácidos orgánicos o las pérdidas por volatilización y la

liberación de compuestos amoniacales; finalmente, el pH disminuye ligeramente durante la

fase de enfriamiento y maduración a valores entre 7 y 8 debido a la formación de compuestos

húmicos con efecto tampón (Iglesias Jiménez and Pérez García, 1991). De manera que, el

seguimiento del pH informa de los procesos de degradación, de manera que pH ácidos

pueden ser indicativos de anaerobiosis o de que el material no está suficientemente maduro.

Sin embargo, el pH no suele considerarse como un indicador de estabilidad ya que existen

Introducción

16

excepciones a la curva estándar y, por ejemplo, residuos ácidos pueden generar compost

ácidos aunque las condiciones de proceso sean óptimas (Epstein, 1996).

Materia orgánica y nutrientes

A lo largo del proceso de compostaje la materia orgánica desciende debido a su

mineralización y a la pérdida de carbono en forma de dióxido de carbono como resultado de

la respiración microbiana, sobre todo, en la fase biooxidativa. Mientras que, en la fase de

maduración prevalecen los procesos de síntesis de compuestos húmicos frente a los procesos

de mineralización. A medida que el proceso de compostaje avanza, el contenido de carbono se

reduce y la producción de dióxido de carbono decrece. La actividad respiratoria bien

mediante la medición del dióxido de carbono producido o el oxígeno consumido por los

microorganismos heterótrofos aerobios son indicadores de la evolución de la actividad

biológica y, consecuentemente, de la estabilidad de un compost. La actividad respiratoria

también se puede medir indirectamente mediante la determinación del calor liberado, por

ejemplo, mediante el test de autocalentamiento.

El nitrógeno orgánico está fácilmente disponible cuando se encuentra en las formas

proteínicas, peptídicas o en forma de aminoácidos libres. Por otro lado, debido a la resistencia

de las moléculas de quitina y lignina al ataque microbiano, la pequeña cantidad de nitrógeno

en ellas se libera lentamente (Diaz et al., 2002). Así, el nitrógeno orgánico es mineralizado a

amoníaco mediante procesos de amonificación. En disolución el amonio se puede o bien,

transformar en nuevos compuestos orgánicos por la microbiota o bien, transformarse en

nitrato mediante bacterias nitrificantes que presentan valores óptimos de temperatura por

debajo de 40ºC y condiciones aerobias. Además, a pH altos y temperaturas elevadas el

amoníaco puede volatilizarse. Por lo general, el nitrógeno decrece a lo largo del compostaje,

el amonio desciende y el nitrato aumenta, de manera que, la evolución del ratio

amonio/nitrato proporciona información sobre cómo se ha llevado a cabo el proceso de

compostaje y puede ser empleado como indicador de madurez.

Durante el compostaje, la transformación del sustrato está condicionada por la

naturaleza del material orgánico de acuerdo con su degradabilidad (Haug, 1993). Los

almidones, azúcares y grasas se descomponen o mineralizan mucho más rápido que las

proteínas o celulosa, mientras que la lignina es muy resistente a la mineralización (Epstein,

2011). Barrington et al. (2002) estudiaron las pérdidas de carbono y nitrógeno tras el

compostaje de agentes estructurantes observando que las pérdidas de carbono variaron

entre el 14-52% y las pérdidas de nitrógeno entre el 38-69%. Bernal et al. (2009) mostraron

Introducción

17

las pérdidas de carbono y nitrógeno de distintos procesos de compostaje de estiércoles,

observando pérdidas de carbono de hasta 72% y de nitrógeno del 60% durante el compostaje

de estiércol de cerdo. Este estudio mostró que el sistema y las condiciones de compostaje, las

características tanto del material de cama como del agente estructurante añadido para el

compostaje e incluso las condiciones ambientales de la estación del año ejercen una gran

influencia sobre la mineralización de la materia orgánica durante el compostaje.

Durante la fase de maduración los compuestos más resistentes a los ataques

microbianos y parcialmente transformados, participan en la formación de los compuestos de

naturaleza húmica. Las fracciones húmicas condicionan la calidad del compost final obtenido

y determinan el grado de estabilización y madurez de la materia orgánica del compost. (Sequi

et al., 1986) propusieron un índice de humificación, calculado como el cociente entre la

fracción de carbono orgánico no humificado y el de la fracción humificada, que se basa en que

los materiales no fenólicos (no humificados) se descomponen en materiales polifenólicos

(humificados). En la actualidad se utilizan técnicas que ofrecen un conocimiento más

detallado de la composición y estructura de la fracción húmica de la materia orgánica como

técnicas electroforéticas, técnicas cromatográficas, espectroscopia IR y FTIR, técnicas de

resonancia magnética nuclear, etc. El estudio del progreso de la humificación mediante

diferentes índices de humificación puede ofrecer una valiosa información acerca de la

transformación de la materia orgánica durante el compostaje (Ciavatta et al., 1993; Sequi et

al., 1986).

2.4. Microorganismos implicados en el compostaje

En condiciones aerobias el principal factor selectivo de las poblaciones microbianas es

la temperatura que determina la tasa de las actividades metabólicas y define también las

fases que comúnmente se desarrollan durante el compostaje (Ryckeboer et al., 2003) y que

fueron expuestas en el apartado anterior. Así mismo, la naturaleza de los sustratos orgánicos

es también un factor importante para determinar la dinámica y la diversidad microbianas

durante el compostaje (Klammer et al., 2008; Ryckeboer et al., 2003; Vargas-García et al.,

2010). Los residuos orgánicos poseen diferentes nutrientes en distintas formas disponibles y

que han sido obtenidos, almacenados o tratados mediante procesos que pueden provocar

cambios en sus poblaciones microbianas. Al comienzo del proceso de compostaje, la adición y

mezcla con distintos sustratos colaboran en incorporar microbiota y diferentes fuentes de

nutrientes, más o menos disponibles, así como, modificaciones físicas que también afectan a

la microbiota. Al inicio del proceso se produce la colonización por parte de los

microorganismos que son capaces de alimentarse de los nutrientes disponibles, en las formas

Introducción

18

en las que se encuentran y que se van a desarrollar de manera prolífica, siendo estos

microorganismos principalmente bacterias. Esto se debe a que las bacterias presentan mayor

ventaja competitiva que los hongos ya que utilizan una amplia gama de enzimas para

degradar químicamente una gran variedad de compuestos orgánicos y su ciclo de

reproducción y desarrollo es más corto. Las comunidades bacterianas se adaptan a los

cambios que se suceden rápidamente en la disponibilidad de sustrato y a otros factores como

la humedad y temperatura. En consecuencia, las bacterias son responsables de la mayor parte

de la descomposición inicial y la generación de calor durante el proceso de compostaje,

siempre que se cumplan los principales requisitos de crecimiento como humedad, pH,

aireación, etc. (Ryckeboer et al., 2003). El otro grupo representativo son los hongos que

presentan una menor tolerancia a las altas temperaturas y un crecimiento más lento, por lo

que son desplazados por las bacterias en condiciones termófilas. Sin embargo, en las fases

finales los hongos, como consecuencia del descenso de humedad, toman ventaja. Cuando las

temperaturas exceden los valores por encima de los 50-60ºC se produce una selección

favorable para las bacterias frente a los hongos, en especial, hacia las bacterias Gram-

positivas del género Bacillus. Por encima de valores superiores a 70ºC se produce la muerte

de la mayoría de las bacterias pudiendo permanecer en el sustrato los hongos en forma

esporulada y especies de bacterias Gram-negativas Thermus e microorganismos del reino

Archaea. Strom (1985) mostró que la temperatura máxima deseable para el compostaje, de

manera que se mantenga una alta diversidad de especies para la estabilidad de la población y

la versatilidad metabólica, es de 60ºC. Dentro de las bacterias Gram-positivas cabe destacar

las actinobacterias que son termotolerantes y termofílicas, crecen mejor cuando el sustrato

está húmedo y correctamente oxigenado y presentan preferencias por pH neutros y alcalinos.

También es destacable que en condiciones termófilas, con valores entorno a los 45-55ºC, y

una humedad baja, puede aumentar la presencia de hongos termófilos como Aspergillus sp. y

Mucor sp. Los hongos son, en general, más importantes para la degradación de la celulosa que

las bacterias, lo cual es especialmente significativo cuando la celulosa está incrustada con

lignina, por ejemplo, en madera o paja (Insam and de Bertoldi, 2007). Durante la fase de

enfriamiento o maduración de los compost, se observa una gran diversidad metabólica de

bacterias mesófilas que juegan una función esencial para la estabilización del compost (Beffa

et al., 1996).

El estudio de la microbiología del proceso de compostaje se realiza desde los años 50-

60 y, debido a su complejidad, continúa en la actualidad. (Ryckeboer et al., 2003) presentaron

un estudio detallado sobre las bacterias y hongos identificados en distintos trabajos de

Introducción

19

investigación durante el compostaje de diferentes residuos. Así en la Tabla 3 se presenta un

resumen de algunas de los géneros más representativos:

Tabla 3. Ejemplos de organismos identificados durante el compostaje (Ryckeboer et al., 2003).

Género/especie Tipología Fases predominantes

Bacillus sp. Bacteria Gram+ Termofílica y mesofílica

Brevibacillus sp. Bacteria Gram+ Termofílica y mesofílica

Clostridium sp. Bacteria Gram+ Termofílica y mesofílica

Enterobacter sp. Bacteria Gram- Termofílica y mesofílica

Flavobacterium sp. Bacteria Gram- mesofílica

Methylobacterium sp. Bacteria Gram- mesofílica

Micromonospora sp. Bacteria Gram+ Termofílica

Nocardia sp. Bacteria Gram+ Termofílica y mesofílica

Paenibacillus sp. Bacteria Gram+ Termofílica y mesofílica

Pseudomonas sp. Bacteria Gram- mesofílica

Serratia sp. Bacteria Gram- mesofílica

Streptomyces sp. Bacteria Gram+ Termofílica y mesofílica

Thermoactinomyces sp. Bacteria Gram+ Termofílica

Thermopolyspora sp Bacteria Gram+ Termofílica

Thermus sp. Bacteria Gram- Termofílica

Acremonium sp. Hongos Mesofílica

Aspergillus sp. Hongos Mesofílica

Aspergillus fumigatus Hongos Termofílica

Chaetomium thermophile Hongos Termofílica

Coprinus sp. Hongos Mesofílica

Doratomyces sp. Hongos Mesofílica

Fusarium sp. Hongos Mesofílica

Gliocladium sp. Hongos Mesofílica

Humicola sp. Hongos Termofílica y mesofílica

Mortierella sp. Hongos Mesofílica

Mucor sp. Hongos Mesofílica

Paecilomyces sp. Hongos Mesofílica

Penicillium sp. Hongos Mesofílica

Scopulariopsis sp. Hongos Mesofílica

Talaromyces sp. Hongos Termofílica

Thermomyces sp. Hongos Termofílica

Trichoderma sp. Hongos Termofílica

Muchos de los trabajos citados por estos autores están basados en técnicas de cultivo. A

lo largo de los años se han mejorado las técnicas y ampliado los estudios sobre distintos

residuos y distintas tecnologías de proceso. Existen en los últimos 10 años numerosos

artículos que estudian los microorganismos durante el proceso de compostaje existentes

mediante diversas técnicas, tanto innovadoras como conservadoras, que proporcionan

información de perfiles de la comunidad microbiana y/o de especies concretas. En la Tabla 4

se presentan algunos ejemplos.

Introducción

20

Tabla 4. Algunas referencias de investigaciones de la microbiología del proceso de compostaje

en los últimos años.

Residuo Técnica de estudio Sistema y tiempo de proceso

Referencia

Residuo sólido municipal Cultivo Pila volteada

26 semanas

(Rebollido et

al., 2008)

Lodo de refinería de aceite

vegetal y basura doméstica Análisis de PLFAs

Pila volteada

5 meses

(Amir et al.,

2010)

Biorresiduos municipales e

industriales

Análisis de PLFAs

Extracción de ADN y análisis

PCR con clonación y

secuenciación

2 tambores de

160m3 3 túneles de

150m3

Tambor a escala

piloto 5m3

(Hultman et

al., 2010)

Residuos hortícolas, lodo de

depuradora y residuos

urbanos

Cultivo Pilas volteadas

1.5m3 185-125 días

(Vargas-

García et al.,

2010)

Residuos de aceitunas

Aislamiento en medio de

cultivo e identificación

mediante 16S sRNA

Extracción de ADN y análisis

PCR-DGGE

Pila volteada 3.5m3

140 días

(Federici et

al., 2011)

Plantas de tomate sin fruto

y virutas de pino

Aislamiento en medio de

cultivo y secuenciación parcial

16S rRNA

Pilas con aireación

forzada y volteos

189 días

(López-

González et

al., 2015)

Estiércoles. Tallos de

tomate y residuos de col.

Residuos verde.

Desperdicios de cocina

Residuos sólidos urbanos

Extracción de ADN y análisis

PCR-DGGE

Pilas 3m3

40 días

(Wang et al.,

2015)

Residuos de castaño Extracción de ADN y análisis

PCR-DGGE

Pilas en contenedor

de 4.05m3

345 días

(Ventorino et

al., 2016)

Alperujo y estiércol de

cerdo

Extracción de ADN,

cuantificación

PCR y pirosecuenciación

Pilas volteadas 10t

22 semanas

(Tortosa et

al., 2017)

2.5. Sistemas de compostaje

De manera general, la elección de un sistema o tecnología de compostaje depende de

varios factores como pueden ser: tipología y cantidad de residuos, consideraciones

económicas, aspectos legales, localización, aspectos ambientales y calidad del producto, entre

otros (Epstein, 2011). En la Tabla 5 se presentan algunos de los sistemas más comunes,

clasificados según su relación con el medio y la mezcla o cambio de posición.

Introducción

21

Tabla 5. Clasificación de los sistemas de compostaje más comúnmente empleados (Diaz et al.,

2007; Epstein, 2011; Moreno and Moral, 2008).

Sistemas abiertos Sistemas Semiabiertos Sistemas cerrados

Estáticos Montones/Pilas/

Mesetas con aireación

pasiva o forzada

Pilas con cubierta

semipermeable aireadas

Contenedores/Túneles

aireados

Dinámicos Pilas volteadas

Mesetas volteadas

Trincheras

Pilas/Mesetas en naves

cerradas volteadas

Túneles dinámicos

Tambores

Diaz et al. (2007) establecieron una diferenciación de los sistemas de compostaje en

dos grandes grupos “windrow”, que hace referencia al acumulo del material a compostar en

pilas o hileras más o menos alargadas, e “in vessel” donde el material se dispone confinado

dentro de un reactor.

Desde un punto de vista operativo, es importante destacar que en muchos casos se

considera que el proceso de compostaje se puede diferenciar únicamente en dos etapas. Una

etapa biooxidativa o intensiva que se corresponde principalmente con las 2 primeras fases

del proceso de compostaje, mesófila inicial y termófila, que se caracteriza por altas

temperaturas, elevado consumo de oxígeno y la producción de emisiones gaseosas y líquidas

(Haug, 1993). Esta fase biooxidativa está condicionada por los procesos de descomposición

de la materia orgánica y, por lo general, en las plantas industriales se desarrolla con una

tecnología concreta y cuya duración va desde los pocos días hasta meses. La segunda fase se

corresponde esencialmente con la etapa de enfriamiento y maduración. Suele tener una

duración superior a la fase biooxidativa aunque el tiempo de residencia en esta fase está

condicionado por las características del material de partida y las condiciones

medioambientales y de operación de la planta de compostaje (Diaz et al., 2002). En general,

presenta un menor control de proceso y, en muchos casos, se limita erróneamente al

almacenamiento del compost para su ensacado (Rynk, 2000). Así una fase biooxidativa corta

y/o en un material energético puede reactivarse en su estancia en la zona de maduración, por

lo que la falta de control puede conllevar problemas de anaerobiosis, lixiviación, olores,

ralentización del proceso y descenso de la calidad del compost (formación de metabolitos

secundarios no deseables y compost con mal olor). La fase biooxidativa se realiza mediante

alguna de las tecnologías de la Tabla 5, mientras que la estancia en maduración se lleva a cabo

en pilas o mesetas, con mayor o menor intervención dependiendo de la idiosincrasia de la

planta de tratamiento o del espacio disponible. Particularmente, la mayoría de las plantas de

Introducción

22

compostaje que emplean como tecnología los sistemas “in vessel” implican el uso de mesetas

para madurar el compost (Diaz et al., 2007).

De manera resumida, se exponen algunas de las características principales de los

sistemas de compostaje:

Las pilas o mesetas pueden desarrollarse con mayor o menor tecnología, en

condiciones más o menos protegidas del exterior y diferenciándose principalmente según su

método de aireación: volteado y aireación forzada o estática. La manera más sencilla consiste

en la formación de una pila y su mantenimiento en condiciones estáticas mediante la

realización de una buena mezcla inicial con el material estructurante y una forma de la pila

adecuada que faciliten la aireación pasiva. Es una tecnología de bajo coste pero puede

presentar problemas para asegurar la reducción de los patógenos por debajo de los límites

legales al no producirse la mezcla del material. Las pilas estáticas también pueden airearse de

manera forzada mediante dos configuraciones diferenciadas: modelo Betsville y modelo

Rutgers. El modelo Betsville fue el primero en desarrollarse y consiste en la colocación de

tuberías por el suelo y encima la mezcla a compostar. Se genera presión negativa en la masa a

compostar mediante succión con un ventilador centrífugo según concentración de oxígeno,

entrando el aire exterior en la pila y recogiéndose el aire del interior del material en un

biofiltro para reducir los olores. En el modelo Rutgers el aire es forzado a pasar a través de la

pila mediante presión positiva lo cual permite garantizar la oxigenación y mayor control de la

temperatura de la pila según las necesidades de proceso. En ambos sistemas la formación de

la mezcla es esencial requiriéndose unas condiciones de porosidad y humedad que permitan

crear las condiciones adecuadas para que la aireación mantenga la aerobiosis en la pila. El

sistema de pila aireada puede ser abierto o semiabierto mediante el uso de lonas o la

realización de las pilas bajo cubierta o con muros. Las mesetas o pilas volteadas son uno de

los sistemas más ampliamente utilizados debido a su baja inversión y a su aplicabilidad en

una gran variedad de residuos. El volteado consiste en la destrucción-reconstrucción de la

pila y en la mayoría de las plantas se emplean palas o volteadoras mediante las cuales el

material se agita y/o mezcla volviendo a depositar el material en una forma semejante a la de

inicio (Figura 5). La dimensión de las mesetas o pilas varían en función de la tecnología de

volteo y a la envergadura de la maquinaria empleada. La frecuencia de los volteos es variable

siendo función de las variables de proceso (oxígeno, temperatura, humedad, etc) y las

variables técnicas y económicas de la planta. Al igual que para el sistema estático, pueden ser

abiertos o semiabiertos mediante el uso de lonas o la realización de las pilas bajo cubierta.

Introducción

23

Figura 5. Imagen de una pila de compostaje a la izquierda y detalle de un volteo con pala a la

derecha.

En cuanto a los sistemas “in vessel” destaca el sistema estático en túneles,

contenedores, reactor estático o similar que mantienen la masa a compostar en un sistema

cerrado en el cual la aireación se realiza mediante aireación forzada. Este método requiere

mayor inversión pero presenta un mayor control del proceso: tratamiento de gases, recogida

de lixiviados, sistema de toma de datos de variables básicas mediante sensores en paredes,

riego por aspersión, etc. El tiempo de residencia habitual en el sistema de túneles se sitúa en

torno a los 14 días. Los reactores estáticos son sistemas que permite la monitorización del

proceso y, por lo tanto, se han empleado en investigación experimental (Figura 6). Otro

sistema “in vessel” es el reactor rotatorio que se diferencia de otros sistemas cerrados en que,

además de la posibilidad de introducir la ventilación forzada, el material gira debido a la

propia rotación del reactor, evitando su compactación y disminuyendo su estratificación

gracias a la homogenización provocada por el giro del tambor.

Las trincheras son un sistema semiabierto que se puede considerar tanto “in vessel”

como “in windrow”. Puede combinar la ventilación forzada, mediante la inclusión de suelo

perforado por donde se introduce el aire, y el sistema de volteos a lo largo de la trinchera

mediante el empleo de volteadoras tipo puente. Esta volteadora circula sobre raíles

dispuestos en los muros longitudinales que delimitan el material (Figura 6).

Introducción

24

Figura 6. A la izquierda, bioreactor de residuos orgánicos (BIORESOR) del Equipo de

Biotecnología Ambiental de la Universidad de Vigo (Patente 009900981) y a la derecha, detalle

de una trinchera de compostaje.

Introducción

25

3. VERMICOMPOSTAJE

3.1. Generalidades

El vermicompostaje es una técnica de degradación de sustratos orgánicos en la cual se

emplean lombrices de tierra para obtener un producto de alto valor agronómico denominado

vermicompost. Esta tecnología utiliza la actividad de algunas especies de lombrices de tierra

que en presencia de residuos orgánicos, directamente mediante su capacidad detritívora y, de

manera indirecta, mediante su interacción con la microbiota, provocan una aceleración en la

descomposición de la materia orgánica. Las lombrices de tierra se alimentan de los residuos,

digiriéndolos y triturándolos y, por lo tanto, provocando cambios físicos, químicos y

biológicos en las propiedades del residuo, debido a la interacción con la propia microbiota

presente en el intestino de las lombrices y porque algunos microorganismos presentes en los

residuos forman parte de su dieta (Edwards and Fletcher, 1988). Domínguez (2004) definió

al vermicompostaje como un proceso biooxidativo en el que los lombrices interaccionan con

los microorganimos y otra fauna descomponedora, acelerando la degradación y estabilización

de la materia orgánica.

3.2. Organismos implicados en el vermicompostaje

En un proceso de vermicompostaje la descomposición de la materia orgánica está

gobernada principalmente por las lombrices de tierra y los microorganismos. Se puede

encontrar otra fauna asociada como cochinillas, tijeretas, larvas de moscas, enquitreidos,

nematodos, ácaros, etc. aunque su efecto en el proceso es menor si éste se desarrolla de

manera controlada y bajo un correcto manejo.

3.2.1. Lombrices de tierra

Las lombrices de tierra son anélidos oligoquetos clitelados que habitan en los suelos y

cuya actividad ejerce un efecto importante en la estructura y en los servicios ecosistémicos

(Blouin et al., 2013). En la actualidad, se considera que existen más de 8000 especies

descritas de Oligochaeta de las cuales la mitad son lombrices terrestres (Reynolds and

Wetzel, 2016) y representan la mayor biomasa animal de la superficie del suelo en la mayoría

de ecosistemas terrestres (Lavelle and Spain, 2001). De acuerdo con Bouché (1977), las

lombrices de tierra ocupan diferentes nichos ecológicos y presentan distintos ciclos vida,

procesos de movimiento y estrategias de alimentación, por lo que se pueden diferenciar en

epigeas, endogeas y anécicas. En la Tabla 6 se presentan las principales características de las

tres categorías ecológicas.

Introducción

26

Tabla 6. Principales características de las tres categorías ecológicas de las lombrices de tierra

(Benckiser, 1997; Coleman et al., 2004; Edwards and Bohlen, 1996; Karaca, 2010).

Epigeas Endogeas Anécicas

Hábitat Horizonte orgánico. Horizonte mineral Horizonte orgánico y

mineral.

Fuente de alimento Hojarasca en

descomposición

Suelo mineral rico en

materia orgánica

Hojarasca en

descomposición

Estrategia alimentación

Detritívora Geófaga Detritívora

Galerías Ninguna

Continuas y

horizontales a 10-15

cm

Verticales desde capa

mineral a la superficie.

Grandes y permanentes

Musculatura Poco desarrollada Desarrollada Bien desarrollada

Tasa de reproducción Alta Limitada Limitada

Tamaño adultos Pequeño Mediano Grande

Pigmentación Uniforme e intensa Ausente o levemente

pigmentada

Dorsalmente y anterior.

Moderada

Movilidad Rápida en respuesta a

perturbaciones Lenta

Moderada en las

galerías

Condiciones adversas (sequedad)

Estado de capullo Diapausa Inactividad

Ejemplos Eisenia fetida,

Eisenia andrei

Dendrobaena octaedra

Aporrectodea

calaginosa Allobophora

chlototica

Lumbricus terrestris

Apporeoctodea longa

De la totalidad de lombrices descritas sólo algunas de ellas se emplean actualmente en

vermicompostaje, siendo la mayoría de ellas pertenecientes a la clase ecológica epigea. Ello se

debe a que las lombrices deben cumplir una serie de requisitos para que su cultivo y su uso

controlado en el proceso de vermicompostaje sea posible:

− Alta capacidad de colonización de distintos residuos orgánicos.

− Elevada capacidad de consumir y digerir sustratos orgánicos.

− Alta capacidad reproductiva y viabilidad de capullos.

− Rápido crecimiento y ciclo de vida corto.

− Rango de tolerancia a variaciones ambientales amplio.

− Elevada resistencia a la manipulación.

Las especies de lombrices de tierra más frecuentemente empleadas en procesos de

vermicompostaje son Eisenia andrei y Eisenia fetida. Se consideraron la misma especie

debido a que comparten parecido morfológico, ciclos de vida similares y forman colonias

mixtas en pilas de estiércol (Bouché, 1972). Sin embargo, se han demostrado sus diferencias

biológicas, reproductivas y filogenéticas (Domínguez et al., 2005; Pérez-Losada et al., 2005) y

Introducción

27

se ha constatado que bajo condiciones controladas, es decir, en procesos de vermicompostaje,

E. andrei puede llegar a desplazar a E. fetida, debido a su mayor tasa de crecimiento y

reproducción. Ambas se caracterizan por poseer un ciclo de vida más corto en comparación

con otras especies de lombrices de tierra empleadas en procesos de vermicompostaje (Figura

7). Así mismo, consumen importantes cantidades de residuos, toleran un amplio rango de

temperatura y humedad y son fácilmente manejables (Domínguez and Edwards, 2011). Se

conocen habitualmente con el nombre de lombriz roja californiana (E. andrei), por su

pigmentación homocrómica rojo oscuro, y lombriz tigre (E. fetida), debido a su color marrón

con bandas intersegmentarias amarillentas. Ambas son especies de climas templados aunque

se han convertido en especies cosmopolitas tanto por su uso en vermicompostaje y

vermicultura como por su empleo en técnicas de ecotoxicología (Domínguez and Edwards,

2011; Voua Otomo et al., 2013).

Figura 7. Ciclo de vida de Eisenia andrei y Eisenia fetida. Modificado de Domínguez and Pérez-

Díaz (2011)

Otras especies de climas templados empleadas en vermicompostaje son Dendrobaena

venetta, Dendrodrilus rubidus y Lumbricus rubellus. Esta última, presenta gran interés en

vermicultura debido a su tamaño, lo cual la hace útil como cebo de pesca y en producción de

harina de lombriz. En cuanto a las especies de lombriz de clima tropical, las más empleadas

en vermicompostaje son Eudrilus eugeniae y Perionyx excavatus. Ambas requieren que el

vermicompostaje se desarrolle bajo control de la temperatura debido a sus estrechos

márgenes de tolerancia a la temperatura ambiental. E. eugeniae es la lombriz de tierra más

grande empleada en procesos de vermicompostaje con un peso medio de 3.5g (Domínguez

and Edwards, 2011; Edwards and Arancon, 2004)

Introducción

28

3.2.2. Microorganismos

Los microorganimos juegan un papel crucial en la descomposición de la materia

orgánica durante el vermicompostaje. En el proceso se ven involucrados una gran cantidad de

microorganismos tanto bacterias, hongos, actinomicetos como algas y protozoos. A diferencia

del compostaje, donde se produce una sucesión microbiana entre mesófilos y termófilos, el

vermicompostaje es un proceso que se desarrolla en condiciones mesofílicas.

Como muchos autores han expuesto anteriormente (Aira et al., 2002; Brown, 1995;

Domínguez, 2011; Lavelle and Spain, 2001) las lombrices de tierra ejercen un efecto directo e

indirecto sobre las comunidades microbianas. En la Figura 8 se exponen las interacciones

ejercidas por las lombrices sobre la microbiota.

Figura 8. Esquema de las interacciones de las lombrices de tierra sobre los microorganismos

(elaboración propia)

Las lombrices de tierra se alimentan de los sustratos orgánicos que, al pasar a través de

la molleja, se fragmentan, aumentando la relación superficie/volumen de las partículas y

falicitando a los microorganimos el acceso al alimento y, por lo tanto, a su colonización. Así

mismo, el alimento ingerido por las lombrices sufre procesos de degradación enzimática a

causa de la actividad de las propias enzimas segregadas en el mucus y por la pared intestinal

de las lombrices, por la actividad de las enzimas segregadas por los microorganismos

endosimbiontes del intestino y por la actividad de la microbiota ingerida propia del sustrato.

De esta manera, el material orgánico digerido se dispersa en forma de excrementos, o casts,

en coexistencia con el sustrato no ingerido. Estos casts recién excretados contienen

Introducción

29

nutrientes y microbiota diferente al material previamente ingerido y son dispersados por el

sustrato al desplazarse las lombrices. Además de los casts, las lombrices de tierra secretan al

exterior mucus y otras sustancias como urea y amonio que constituyen un aporte de

nutrientes fácilmente asimilables para los microorganimos (Domínguez et al., 2011). La

microbiota del residuo forma parte de la dieta de las lombrices, de manera que se ha

demostrado que se alimentan principalmente de hongos, tanto de manera selectiva como

generalista. Algunas especies de bacterias pueden ser digeridas, otras pueden sobrevivir al

paso a través del tracto intestinal y otras bacterias pueden crecer y volverse más activas.

También pueden constituir parte de la dieta de las lombrices de tierra los protozoos, algas y

nematodos, bien mediante una ingesta accidental o por preferencia. La cantidad de

microorganismos consumidos y la capacidad de las lombrices para digerirlos varían con la

especie de lombriz de tierra, su categoría ecológica, el tipo de alimento y las condiciones

ambientales, de ahí que el papel de los microorganismos como fuente de nutrición de las

lombrices no esté claro (Brown and Doube, 2004; Curry and Schmidt, 2007). La actividad de

las lombrices a través del residuo provoca modificaciones físicas, como consecuencia de su

movimiento y excavación, como son la aireación y homogeneización del sustrato, lo cua,

favorece la actividad microbiana y mejora la descomposición del residuo (Domínguez, 2004).

Además, existe una competencia entre las lombrices y la microbiota por los recursos del

sustrato. Tiunov and Scheu (2004) observaron que la presencia de lombrices incrementaba la

limitación en carbono para los microorganismos del suelo a causa del aumento de nitrógeno y

fósforo en los casts o a un agotamiento directo de los recursos de carbono fácilmente

disponibles.

En la Tabla 7 se presentan investigaciones que estudian las comunidades microbianas

durante el vermicompostaje o en el vermicompost.

Introducción

30

Tabla 7. Referencias de investigaciones de la microbiología del proceso de vermicompostaje.

Residuo/ sistema Lombriz de tierra Técnica de estudio Referencia

Residuo fresco de aceituna y

estiércol de oveja. Camas 1.5 m2

E. fetida

5g/100 g ps

Extracción ADN y análisis

PCR-DGGE y PCR tiempo

real

(Vivas et al.,

2009)

Purín de cerdo. Módulos 1,5kg o

3kg. Recipientes 200g

E. fetida

500 indiv/módulo

50indiv/recipiente

Coliformes por filtración y

cultivo

(Monroy et al.,

2009)

Residuo de la industria azucarera.

Recipiente 2,5 kg

E. eugeniae

15g/kg

Extracción ADN y análisis

PCR-DGGE

(Sen and

Chandra,

2009)

Purín de cerdo.

Módulos 1,5kg o 3 kg

E. fetida

500 indiv.

maduros/módulo

Análisis PLFAs

(Gómez-

Brandón et al.,

2011a)

Vermicompost a partir de

distintos residuos y sistemas de

vermicompostaje.

E. fetida Extracción ADN y análisis

PCR-DGGE, Microarray

(Fernández

Gómez et al.,

2012)

Residuos agroindustriales de

vino, aceituna y estiércol. Cajas

2,4m2

E. fetida biomasa

equiv. 3% m.o.

Extracción ADN y análisis

PCR-DGGE y PCR tiempo

real

(Castillo et al.,

2013)

Residuos vegetales.

Reactores 200g

E. fetida

10 g/reactor

Extracción ADN y análisis

PCR-DGGE

(Huang et al.,

2013)

Residuos frutas y verduras.

Contenedores 2kg

E. fetida 100

indiv/contenedor

PCR tiempo real, PCR-

DGGE, pirosecuenciación

(Huang et al.,

2014)

Hojarasca de maíz y excrementos

de pollo. Recipientes 200g

E. fetida

6 indiv/recipiente

Extracción ADN y análisis

PCR, pirosecuenciación

(Chen et al.,

2015b)

3.3. Condiciones previas y control de proceso

El vermicompostaje es un proceso biotecnológico y, como tal, presenta unos

requerimientos iniciales impuestos por la propia biología de la microbiota y la fauna

implicada. Requiere un control del proceso para que los organismos se desarrollen

adecuadamente en condiciones y tiempo que permitan la obtención del producto en apto

para su uso posterior.

3.3.1. Condiciones previas

Como se comentó anteriormente, las lombrices implicadas en procesos de

vermicompostaje presentan diferentes rangos de tolerancia a las condiciones ambientales.

Estos factores deben ser tenidos en cuenta cuando se busca tratar un residuo orgánico. A

continuación se exponen algunos de los parámetros más importantes:

Introducción

31

Humedad

Las lombrices de tierra están protegidas del exterior con una fina cutícula a través de la

cual se efectúa el intercambio gaseoso. Su epidermis cuenta con glándulas mucosas que

excretan distintos fluidos que las mantienen húmedas, de modo que el oxígeno se disuelve en

esta película superficial (Edwards and Bohlen, 1996). Las condiciones de sequedad pueden

provocar la muerte o la migración de las lombrices hacia zonas más favorables, por lo que, el

mantenimiento de condiciones de humedad óptimas permite aumentar la eficacia en el

tratamiento de los residuos. Los requerimientos de humedad varían dependiendo de la

especie de lombriz y del residuo orgánico a tratar, aunque de manera general se establece un

rango entre 50-90%. A modo de ejemplo, Domínguez and Edwards (1997) observaron que E.

andrei crece y madura en estiércol de cerdo entre 65-90% de contenido de humedad siendo

el óptimo 85%.

Temperatura

Las distintas especies de lombriz difieren en cuanto a los rangos tolerables de

temperatura. Loehr et al. (1985) y Neuhauser et al. (1988) observaron que Dendrobaena

veneta, Eisenia fetida, Eudrilus eugeniae, Perionyx excavatus and Pheretima hawayana

presentan temperaturas óptimas para su crecimiento y reproducción en torno a 20-25ºC. A

temperaturas inferiores a 15ºC sólo E. fetida produce capullos mientras que a 30ºC sólo P.

excavatus los produce. Sin embargo, Edwards (1988) observó que a temperaturas inferiores a

9ºC y superiores a 30ºC tanto P. excavatus y E. eugeniae presentaban una mortalidad elevada

siendo su temperatura óptima 25ºC. Estos estudios observaron que las temperaturas óptimas

para la producción de capullos fueron más bajas que las más adecuadas para el crecimiento

de las especies estudiadas. En general, la temperatura óptima para un adecuado proceso de

vermicompostaje se sitúa entre los 10-35ºC.

Estructura física y aireación

Como se ha comentado, la respiración cutánea de las lombrices de tierra exige una

correcta difusión del oxígeno con el medio a través de la pared del cuerpo. Por lo tanto, el

residuo debe poseer una estructura suficientemente porosa para posibilitar concentraciones

de oxígeno óptimas para las lombrices. Condiciones de humedad excesiva, residuos con un

alto contenido graso o compactación del material pueden provocar situaciones de

anaerobiosis afectando al desarrollo de las lombrices por falta de oxígeno y formación de

compuestos tóxicos. Además, la estructura del residuo debe posibilitar el adecuado

movimiento de las lombrices, así mismo, las lombrices, con su propio desplazamiento,

Introducción

32

contribuyen al aumento de la aireación. Otro parámetro que afecta a la porosidad y

estructura es el tamaño de partícula. El residuo debe poseer partículas con superficie

área/volumen accesible a las lombrices. Lowe and Butt (2005) establecieron que el tamaño

de partícula afecta a las tasas de crecimiento y reproducción de lombrices, requiriendo las

lombrices más pequeñas tamaños más reducidos (<1mm).

Concentración de sales y amonio

La mayor parte de la absorción y pérdida de agua en las lombrices se produce a través

de la pared del cuerpo observándose que las lombrices pueden mantener relativamente

constante su presión osmótica interna en soluciones diluidas pero son incapaces de

mantenerla en las más concentradas (Edwards and Bohlen, 1996). Por ello, un exceso de sales

inorgánicas y altos niveles de amonio en el medio resultan altamente tóxicos para las

lombrices inhibiendo su desarrollo y causando la muerte. En el caso de E. fetida y E. andrei se

ha visto que conductividades eléctricas superiores a 8dS m-1 y contenidos de amonio

superiores a 0.5 mg g-1 se consideran tóxicos (Edwards, 1988).

Otros parámetros físico-químicos a considerar son el pH del residuo y sustancias

tóxicas tanto orgánicas como inorgánicas presentes en el sustrato. Las lombrices de tierra

presentan un rango de tolerancia amplio de pH y pueden sobrevivir en condiciones desde

ácidas a alcalinas (pH 5-9) aunque el pH óptimo depende de la especie de lombriz. Por otro

lado, la presencia en el medio de sustancias como son los plaguicidas, antibióticos,

hidrocarburos o metales pesados puede causar efectos adversos en el desarrollo,

supervivencia y actividad de las lombrices, ello dependerá de la dosis de contaminante y la

especie de lombriz (Benitez et al., 1999a; Fernández Gómez et al., 2011; Reinecke et al., 2001;

Thiele-Bruhn, 2003)

La naturaleza y propiedades del residuo orgánico condicionan el desarrollo adecuado

de las lombrices de tierra y, por lo tanto, la eficacia del proceso de vermicompostaje. De esta

manera, en muchos casos es necesario acondicionar el residuo que se pretende tratar de

manera que se posibilite la supervivencia y desarrollo de las lombrices. Para ello se pueden

realizar distintas acciones de manera independiente o en conjunto:

• Mezclas con otros residuos: la finalidad de la mezcla es mejorar las propiedades de los

residuos modificando positivamente su estructura, pH, salinidad, relación C/N, etc.

Por ejemplo, Domínguez et al. (2000) mezclaron lodo de depuradora con diferentes

estructurantes observando mejoras en la reproducción frente al lodo sin estructurar y

Elvira et al. (1998) mejoraron el vermicompostaje de lodos de depuradora de

Introducción

33

industrias del papel y lecherías mediante la adicción de estiércol de vaca. De la misma

manera, Kaushik and Garg (2004) vermicompostaron lodos de la industria textil al

mezclarlos con estiércol de vaca.

• Pre-compostaje: realizar un compostaje previo a la inoculación de las lombrices de

tierra tiene varios objetivos:

− Asegurar la eliminación de patógenos y semillas de malas hierbas al

desarrollarse altas temperaturas durante el compostaje del residuo.

− Eliminar posibles tóxicos para las lombrices de tierra que presentan los

residuos orgánicos frescos.

− En otros casos, los residuos presentan altos contenidos de sustancias

fácilmente asimilables sobre los que puede proliferar de manera importante la

microbiota, aumentando la temperatura e impidiendo la supervivencia de las

lombrices, por lo que una fase previa de pre-compostaje degrada la materia

orgánica más lábil

− Desde una escala industrial, el pre-compostaje facilita el manejo del residuo

para su distribución en los sistemas de vermicompostaje.

Tras un proceso de compostaje puede ser necesario airear el material para enfriarlo o

permitir la liberación de sustancias como el amonio. Diversas investigaciones han

realizado pre-compostaje previamente al vermicompostaje (Tabla 8).

• Otros procesos necesarios para que el residuo pueda ser tratado mediante las

lombrices de tierra incluyen la trituración y la humectación.

Introducción

34

Tabla 8. Referencias de investigaciones que emplean el pre-compostaje previo al

vermicompostaje.

Residuo Tiempo pre-compostaje/Óptimo

Especie de lombriz

Referencia

Residuos de jardín 0, 2, 4, 6 semanas/Limitar al

mínimo E. andrei

(Frederickson et

al., 1997)

Biosólidos y papel 28 días E. fetida (Ndegwa and

Thompson, 2001)

Estiércol de ganado 0, 1, 2, 3, 4, 5

semanas/Limitar al mínimo E. fetida

(Gunadi et al.,

2002)

Residuos de cocina, recortes

de césped, papel triturado

0, 6, 9, 12, 15 días/

mejores resultados 9 días

L. rubellus

E. fetida (Nair et al., 2006)

Estiércol de ganado 15 días E. andrei (Lazcano et al.,

2008)

Residuos industria azucarera 60 días E. eugeniae (Sen and Chandra,

2009)

Estiércol y papel 28 días E. fetida (Mupondi et al.,

2010)

Residuos de tomate y

cáscaras de almendra 63 días

E. fetida

E. andrei

(Fornes et al.,

2012)

Estiércol pato con paja y

zeolita 45 días E. fetida

(Wang et al.,

2014)

A estos parámetros hay que añadir los efectos de la densidad de población de

lombrices. Dependiendo de la especie de lombriz de tierra, la calidad del alimento y la

superficie del sistema de vermicompostaje empleado, existe unos valores óptimos para la

densidad de población. Un exceso de densidad puede provocar competencias por el alimento

mientras que un defecto de individuos puede generar problemas en la reproducción. Del

mismo modo, el objetivo perseguido también puede variar la densidad necesaria, ya que si se

busca reproducción, crecimiento o el tratamiento del residuo, las densidades pueden diferir.

Domínguez and Edwards (1997) encontraron que en condiciones óptimas de humedad 8

individuos de E. andrei en 43.61g de purín (masa seca) era la densidad ideal para el

crecimiento y maduración de esta especie. Ndegwa et al. (2000) determinaron que, variando

la tasa de alimentación, la densidad de 1.60kg lombrices m-2 muestra la mayor bioconversión

de biosólidos en biomasa de lombrices y las mejores producciones de vermicompost.

3.3.2. Fases del proceso

Durante el proceso de vermicompostaje se pueden distinguir de manera general dos

fases (Lazcano et al., 2008):

Introducción

35

Fase activa: caracterizada por la actividad de las lombrices que se desplazan por el

residuo y se alimentan de los sustratos orgánicos digiriéndolos, asimilándolos y excretando

casts, lo cual, provoca cambios en las propiedades fisicoquímicas y microbianas del residuo.

El vermicompost resultante es una mezcla de casts y material no procesado por las lombrices.

La duración de esta fase no es fija y depende de la densidad de población, de las especies

empleadas y de la tasa de ingestión y procesado del residuo (Domínguez, 2004)

Fase de maduración: las lombrices se desplazan hacia materiales más frescos, por lo que

los microorganismos asumen el control de la descomposición del material procesado.

Domínguez (2004) explica que los casts están sujetos a un proceso de maduración y a la

acción de los microorganismos y microinvertebrados presentes en el sustrato.

Otra consideración a tener en cuenta del proceso de vermicompostaje es la etapa de

adaptación de las lombrices al residuo. Cuando se dispone en contacto un nuevo residuo con

la población de lombrices que lo va a digerir puede existir un período de baja alimentación o

incluso mortalidad de parte de la población (Elvira et al., 1998; Gajalakshmi et al., 2002;

Haimi and Huhta, 1986). Este período es más habitual en microcosmos experimentales o

sistemas de vermicompostaje de pequeño volumen debido a que las lombrices no encuentran

hábitats ecológicos adecuados y pueden sufrir estrés (Domínguez, 2004). La etapa de

adaptación se puede minimizar de diferentes maneras; comenzando directamente el

vermicompostaje con las lombrices en el propio sustrato de maternidad o cultivo y añadiendo

capas de residuo sobre este sustrato; aclimatando las lombrices mediante la incorporación de

dosis bajas de residuo mezclado con el alimento durante la elaboración del stock de

lombrices; emplear sustratos de escape, como la vermiculita, de manera que las lombrices se

vayan alimentando lentamente del residuo y puedan migrar hacia el sustrato en caso de

requerirlo.

3.3.3. Cambios y control del proceso de vermicompostaje

El vermicompostaje es una tecnología, en general, sencilla y de bajo coste que

aprovecha la capacidad de las lombrices de digerir sustratos orgánicos para bioestabilizar un

residuo orgánico transformándolo en vermicompost. Este proceso, a priori sencillo, tiene

unos requerimientos iniciales, ya presentados en el epígrafe 3.3.1, y también conlleva,

necesariamente, un control del proceso, de manera que las lombrices se desarrollen

adecuadamente y el proceso sea efectivo. A continuación se presentan los parámetros de

seguimiento más ampliamente utilizados:

Introducción

36

Parámetros fisicoquímicos

Condiciones aerobias: la propia actividad de las lombrices posibilita el mantenimiento

de los niveles de oxígeno adecuados a lo largo del proceso de vermicompostaje. Sin embargo,

durante la adición de residuo en sistemas de vermicompostaje de alimentación continua debe

asegurarse que la capa dispuesta presenta un grosor óptimo para evitar fermentaciones y,

con ello, todos los problemas asociados (metabolitos indeseables, anaerobiosis) para las

lombrices de tierra.

pH: en general, el vermicompostaje cambia el pH hacia la neutralidad, con

independencia del pH inicial de los sustratos orgánicos (Pramanik et al., 2007). Si bien, el pH

desciende levemente durante el vermicompostaje como consecuencia de la mineralización

del nitrógeno a nitratos, la formación de ortofosfatos, y la mineralización de materia orgánica

a dióxido de carbono y a otros compuestos orgánicos ácidos intermedios. Singh et al. (2005)

observaron un aumento en el pH en la fase inicial del proceso de vermicompostaje que

atribuyeron al metabolismo microbiano aerobio que tiene como resultado la formación de

hidróxidos básicos en el sustrato en descomposición. Ndegwa et al. (2000) observaron que

distintos sustratos pueden producir diferentes especies intermedias, por lo que, diferentes

residuos muestran un comportamiento distinto en la evolución del pH. Así, autores como Vig

et al. (2011) y Bhat et al. (2013) observaron aumentos en el pH de varios vermicomposts de

distinto origen sugiriendo que podría ser causado por el exceso de nitrógeno orgánico no

requerido por los microorganismos que se libera como amoníaco, se disuelve en agua e

incrementa el pH del vermicompost.

Carbono: durante el vermicompostaje se produce la mineralización del carbono y su

pérdida en forma de dióxido de carbono. Las lombrices y los microorganismos emplean el

carbono como fuente de energía y en su crecimiento celular. Las tasas de reducción del

carbono dependen del residuo, la especie de lombriz, su densidad y el tipo de

vermicompostaje realizado. En experimentos en macetas con E. fetida durante 100 días, Garg

et al. (2006) observaron reducciones del carbono orgánico total (COT) superiores al 50% en

agroresiduos y residuos de cocina, en torno al 40% en residuos municipales y 32% de

reducción en residuo de la industria textil. Elvira et al. (1998) encontraron reducciones del

COT entre 20-43% durante el vermicompostaje en literas de distintos residuos con E. andrei.

Se han visto reducciones en torno al 25% y 31% del COT durante el vermicompostaje en

contenedores de lodo de depuradora con E. fetida en 90 días (Li et al., 2011) y con P.

excavatus en 45 días (Khwairakpam and Bhargava, 2009), respectivamente. Durante el

Introducción

37

vermicompostaje de estiércol de vaca y alperujo con E. andrei, Plaza et al. (2008) obtuvieron

una reducción del 58% del COT en 8 meses.

Nitrógeno: las variaciones en la concentración de nitrógeno en el proceso de

vermicompostaje son variables observándose tanto descensos como aumentos que dependen

del residuo y de las particularidades del proceso. Las lombrices mejoran la mineralización del

nitrógeno y lo consumen empleándolo en su construcción celular. Por otro lado, las lombrices

enriquecen el residuo con nitrógeno mediante la excreción de moco, enzimas y por la propia

descomposición de lombrices muertas. Además, el descenso en contenidos de carbono

orgánico durante el vermicompostaje permite un aumento de la concentración de nitrógeno

(Suthar, 2007). Cambios en el pH hacia condiciones ácidas pueden ser también un factor

importante en la retención de nitrógeno ya que en condiciones alcalinas el amoníaco es

volátil (Hartenstein and Hartenstein, 1981). Por lo tanto, el contenido de nitrógeno en el

vermicompostaje depende del nitrógeno inicial en el residuo y de los procesos de asimilación

y descomposición por parte de las lombrices.

Otros macronutrientes: en general, la concentración de nutrientes aumenta durante el

proceso de vermicompostaje como consecuencia de la mineralización de la materia orgánica

debido a un efecto de concentración. Autores como Elvira et al. (1996), Garg and Kaushik

(2005) o Yadav and Garg (2009) observaron un aumento en el contenido de fósforo, potasio y

calcio tras el vermicompostaje de diferentes residuos. Sin embargo, Orozco et al. (1996) y

Nogales et al. (2005) observaron un descenso en el potasio durante el vermicompostaje de

pulpa de café y residuos vinícolas, respectivamente, debido a su lixiviación por exceso de

agua. Benitez et al. (1996) encontraron que los lixiviados recogidos durante el proceso de

vermicompostaje presentan concentraciones elevadas de potasio.

Metales pesados: los metales pesados tales como Cd, Cu, Ni, Pb, Zn o Hg tienden a

aumentar su concentración total como consecuencia de la mineralización de la materia

orgánica, sin embargo, se ha observado que las formas disponibles tiende a disminuir. Por

ejemplo, Elvira et al. (1995) observaron un descenso en la capacidad de extracción de Mn, Cu

y Zn durante el vermicompostaje de lodos industriales. Singh and Kalamdhad (2013)

observaron que el vermicompostaje era muy eficaz para reducir las fracciones más

biodisponibles de los metales pesados, y que la biodisponibilidad dependía de propiedades

fisicoquímicas y biológicas del vermicompost, siendo el pH uno de los principales factores. El

vermicompostaje tiende a estabilizar los metales pesados redistribuyéndolos desde estados

relativamente lábiles a más inmovilizados (Morgan, 2011), de manera que, tienden a formar

complejos con los ácidos húmicos y las fracciones orgánicas más polimerizadas (Domínguez,

Introducción

38

2004). A este cambio en la disponibilidad de los metales pesados a lo largo del proceso, hay

que añadir que las lombrices de tierra son capaces de acumular un número esencial y no

esencial de metales en sus tejidos. La dinámica de los metales pesados durante el proceso de

vermicompostaje es complejo, posiblemente debido a especificidades del metal y las

interacciones con la matriz (Morgan, 2011).

Humificación: en diversas investigaciones se han observado descensos en contenidos

de carbono solubles y totales y aumentos en sustancias húmicas y , por lo tanto, en las tasas

de humificación tras el vermicompostaje de distintos residuos orgánicos (Elvira et al., 1998,

1996; Plaza et al., 2008; Romero et al., 2007). Domínguez (2004) expuso que las lombrices de

tierra aceleran y mejoran el proceso de humificación debido a la fragmentación de la materia

orgánica, al aumento de la actividad microbiana dentro del intestino de las lombrices y a la

aireación y los volteos del material como consecuencia de la alimentación y el movimiento de

las lombrices.

Parámetros biológicos

Población y estado de las lombrices: tanto la densidad de lombrices como su estado debe

ser controlado a lo largo del proceso de vermicompostaje prestando atención a si las

lombrices se alimentan del residuo, crecen y se reproducen, así como, a la viabilidad de los

nuevos individuos. Cuando se pretende vermicompostar un nuevo residuo es conveniente

conocer sus propiedades, de manera que, se cumplan los requerimientos para el

vermicompostaje (pH, humedad, amonio,…), pero también, realizar estudios de supervivencia

de lombrices de tierra con la especie de interés y la densidad de población que se va a utilizar

posteriormente en el proceso. Al inicio del vermicompostaje pueden producirse mortalidades

elevadas de la población de lombrices causadas por su adaptación a la matriz y a las

condiciones del proceso, así mismo, a medida que el residuo se consume puede aumentar la

mortalidad por falta de alimento. Elvira et al. (1997) mostraron mortalidades altas en

distintas mezclas con lodo de depuradora de una industria de fabricación de papel,

mostrando que en varias de las mezclas la mortalidad no ocurrió por falta de alimento sino

por cambios en las características del sustrato al producirse la degradación de la materia

orgánica.

Las lombrices de tierra alcanzan su madurez sexual después de un período que varía

según la especie de lombriz y las condiciones ambientales. El estudio de la presencia de

individuos tanto clitelados como inmaduros, así como, juveniles y capullos, ofrece

información sobre si las condiciones del residuo y del proceso son favorables. Se ha visto que

Introducción

39

la tasa de producción de capullos está relacionada con la calidad del residuo que es uno de los

factores que determinan el tiempo necesario para alcanzar la madurez sexual y para el inicio

de la reproducción (Edwards et al., 1998).

Actividades microbianas: el análisis de la actividad microbiana aporta información

sobre la degradación de la materia orgánica. Estas actividades se miden, principalmente,

mediante técnicas respiratorias y la determinación de actividades enzimáticas. Las

actividades respiratorias se relacionan con el metabolismo microbiano, de manera que, en la

fase inicial del vermicompostaje, cuando se degradan los compuestos fácilmente asimilables,

la tasa respiratoria es alta y desciende a lo largo del proceso de vermicompostaje como

consecuencia de la reducción de la disponibilidad de recursos para las comunidades

microbianas (Domínguez, 2011). Tanto los microorganismos del residuo como los

endosimbiontes del intestino de las lombrices, producen enzimas tanto intracelulares como

extracelulares que provocan la descomposición de los distintos sustratos orgánicos.

Numerosos autores han estudiado las actividades enzimáticas durante el vermicompostaje de

distintos residuos: purín de cerdo (Aira et al., 2007a, 2007b, 2006), estiércol de ganado

(Lazcano et al., 2008), lodo de depuradora (Benitez et al., 1999b), residuos de aceituna

(Benitez et al., 2005; Melgar et al., 2009; Vivas et al., 2009), residuos vitivinícolas (Nogales et

al., 2005), residuos vegetales de invernadero (Fernández Gómez et al., 2013, 2010a, 2010b),

residuos de la industria azucarera y cenizas (Pramanik and Chung, 2011), residuos

agroindustriales lignocelulósicos (Castillo et al., 2013), residuos de maíz (Chen et al., 2015a),

lodo de panadería (Yadav et al., 2015), residuos de curtiduría (Ravindran et al., 2014), etc. Las

principales enzimas analizadas son la actividad deshidrogenasa y las actividades hidrolíticas

como la β-glucosidasa, proteasa, fosfatasas, celulasas y ureasas, entre otras. Al igual que para

la actividad respiratoria, las fases iniciales presentan, en general, mayor actividad enzimática

reduciéndose a medida que avanza el vermicompostaje. Sin embargo, la interpretación de los

datos enzimáticos es complicado ya que las actividades dependen de numerosos factores y de

la diferente localización de las enzimas que contribuyen a la medida enzimática en el sistema

estudiado (Nannipieri et al., 2002).

Estructura de la comunidad microbiana: es aconsejable estudiar las distintas

poblaciones microbianas durante el vermicompostaje para establecer los efectos que las

lombrices de tierra ejercen sobre los microorganismos, ya que si las lombrices estimulan o

reducen la microbiota, o modifican la estructura y función de las comunidades microbianas,

pueden tener efectos muy diferentes sobre las tasas y forma de descomposición de la materia

orgánica (Domínguez, 2011). Gómez-Brandón et al. (2011b) mostraron que el paso de

Introducción

40

distintos estiércoles a través del intestino de la lombriz E. fetida provocaba la reducción de la

biomasa bacteriana y Fernández Gómez et al. (2010a) observaron que la estructura de las

comunidades fúngicas difería en la etapa de máxima biomasa de lombrices durante el

vermicompostaje de residuos de fruta, sugiriendo ambos trabajos un fuerte efecto asociado al

paso del residuo a través del intestino de las lombrices. Huang et al. (2013) observaron una

reducción de la biomasa microbiana en el vermicompostaje de residuos vegetales pero un

aumento de la diversidad microbiana. Sin embargo, en un estudio posterior, Huang et al.

(2014) mostraron un aumento de las poblaciones bacteriana y fúngica durante el

vermicompostaje lo cual sugiere que los diferentes tipos de sustratos, modos de operación y

condiciones ambientales utilizadas en los experimentos, pueden influir en el crecimiento y

reproducción de lombrices y la microbiota (Gunadi and Edwards, 2003). Además, el tipo de

método empleado para el análisis y seguimiento de la estructura microbiana aporta distinta

información sobre la diversidad y las funciones biólogicas de las distintas comunidades

microbianas implicadas en el vermicompostaje.

Los vermicompost presentan distinta diversidad y carga microbiana que afectarán

positivamente (estímulación de la microbiota del suelo, supresión de patógenos, degradación

de plaguicidas y otros compuestos orgánicos, etc.) o negativamente (inoculación de

patógenos al suelo, exceso de biomasa microbiana y competencia por nutrientes, etc) al

crecimiento de las plantas. Por lo que el estudio de la microbiota es esencial no sólo para la

mejora del proceso de vermicompostaje, sino también para la obtención de un producto de

valor añadido con diferentes usos.

Patógenos: a diferencia del compostaje, durante el vermicompostaje no se alcanzan las

temperaturas que aseguran la eliminación de patógenos, si bien, en diversos residuos se

utiliza el pre-compostaje como acción previa a la inoculación de las lombrices, por lo que este

problema se reduce. En caso del vermicompostaje sobre residuos frescos, se ha visto que las

lombrices reducen la carga patogénica del residuo. Se ha constatado que el vermicompostaje

es efectivo para reducir y eliminar distintos patógenos en biosólidos y lodos (Eastman et al.,

2001; Rodríguez-Canché et al., 2010). Así mismo, Monroy et al. (2009, 2008) observaron

reducciones superiores al 98% de la población de coliformes como resultado del paso de

purín de cerdo a través del intestino en distintas especies de lombriz. Sin embargo, debido a

las diferencias fisicoquímicas y biológicas de los distintos residuos, el sistema de

vermicompostaje y las condiciones de proceso, la reducción de patógenos durante el

vermicompostaje puede diferir. Hénault-Ethier et al. (2016) encontraron que la reducción de

E. coli durante el vermicompostaje de residuos orgánicos separados en origen es causado

Introducción

41

fundamentalmente por la microbiota antagonista, mientras que el efecto de las lombrices

sobre este patógeno es menor. Aira et al. (2011) mostraron que, en un vermirreactor de

alimentación continua con estiércol de vaca, el efecto de las lombrices depende del tipo de

patógeno considerado.

Además, durante el vermicompostaje debe vigilarse la presencia de predadores que

puedan causar daños a las lombrices como aves, ratones, topos, hormigas, etc. El mayor o

menor riesgo vendrá asociado al tipo de sistema de vermicompostaje implantado y al lugar

donde se realice (outdoor or indoor).

3.4. Sistemas de vermicompostaje

Se pueden diferenciar los sistemas de vermicompostaje en función de diferentes

argumentos, por ejemplo:

− Según el sistema de alimentación: discontinua, donde el alimento o residuo orgánico

se añade una única vez y es separado de las lombrices una vez digerido, o continua, el

alimento se añade por tandas, de manera que, cuando una cantidad de residuo es

procesado por las lombrices se añade nuevo residuo.

− Según el nivel de tecnología: tradicional, que hace referencia a sistemas sencillos no

automatizados y de fácil uso, o moderno con sistemas más tecnificados y que

requieren una mayor especialización en el uso.

− Según la escala: doméstica, de un tamaño pequeño para hacer frente a la generación

del residuo orgánico en el hogar; pequeña o mediana escala, para la obtención de

residuo dentro de la propia explotación o instalación de manera que el residuo

orgánico generado se trata y el vermicompost se emplea como fertilizante en la

misma instalación; escala industrial, cuyo objetivo es el tratamiento del residuo y la

producción de vermicompost. Un sistema industrial de vermicompostaje exige que la

planta de procesado este diferenciada en zonas: área de pre-procesamiento o

acondicionamiento del residuo, área de cría o maternidad de lombriz, área de

procesamiento del residuo y área de maduración y almacenamiento del

vermicompost.

A continuación, se presentan, brevemente, los sistemas de vermicompostaje en función

de su tipología, es decir, según los distintos modelos de sistemas empleados o presentes en el

mercado.

Introducción

42

Literas, camas, pilas o zanjas

Es el más similar al proceso natural de degradación, debido a su sencillez y se basa en el

apilamiento del residuo orgánico en espacios más o menos rectangulares o piramidales

(Figura 9). Suelen delimitarse con maderas, ladrillos, bloques, etc., de manera que, se

contenga el residuo y se evite la fuga de las lombrices. Los apilamientos no deben ser

excesivamente altos para evitar el autocalentamiento del residuo y/o compactaciones

excesivas del material.

Figura 9. Foto de una litera tomada de la Junta de Andalucía (izquierda) y litera experimental

del Equipo de Biotecnología Ambiental (derecha)

La alimentación puede ser realizada en una única vez o ser dispuesta en capas

superiores o frontales para que las lombrices avancen horizontal o verticalmente. De manera

habitual, se cubren para protegerlas de las condiciones ambientales y de los depredadores.

Una vez que se alcanza la capacidad de tratamiento es necesario separar el vermicompost de

la población de lombrices, para lo cual se suelen emplear trampas con residuo fresco hacia

donde se muevan las lombrices. Las literas y similares son, en general, sistemas sencillos y

económicos, aunque para la producción de vermicompost a gran escala es necesaria la

utilización de una superficie extensa (Tabla 9).

Tabla 9. Ventajas y desventajas del sistema de vermicompostaje en literas, modificado de

Edwards (2010).

Ventajas Baja inversión económica

Manejo sencillo

Desventajas Mano de obra intensiva

Espacio necesario elevado

Tiempo de proceso lento

Considerables pérdidas de nutrientes por volatilización y

lixiviación

Difícil recoger el vermicompost sin lombrices

Introducción

43

Algunas de estas desventajas se reducen al realizar las pilas bajo techo y en condiciones

ambientales controladas de manera que las condiciones estacionales (temperaturas altas o

bajas, lluvias, insolación, etc.) tengan menor afectación sobre el proceso.

Contenedores

Su uso se emplea tanto a nivel doméstico como industrial. Pueden ser de dos tipos,

modulares y no modulares. Los modulares consisten en el uso de recipientes, bandejas o

similar, apilables y agujereados (Figura 10). Cuando el residuo de un recipiente es consumido

por las lombrices de tierra, se añade residuo fresco en un nuevo recipiente que se deposita

sobre el anterior. Las lombrices pueden moverse a través de los orificios que comunican los

recipientes, accediendo al nuevo recipiente con alimento fresco. El vermicompost se

almacena en los módulos inferiores gracias al desplazamiento vertical de las lombrices,

pudiendo retirarse cuando el vermicompost alcance la madurez deseada, sin interferir en la

continuidad del proceso. Un sistema industrial con contenedores modulares exige una

cantidad importante de módulos, lo cual conlleva requerimientos importantes de personal.

Figura 10. Sistema de vermicompostaje vertical modular doméstico Can-o-worm (izquierda) y

sistema de vermicompostaje en cajas modulares de Ecocelta Galicia S.L.(derecha).

En cuanto a los no modulares, al igual que los anteriores, el residuo se deposita

periódicamente en la superficie del contenedor o recipiente. Las lombrices se mueven en

vertical hacia el alimento fresco dejando el vermicompost en la zona inferior del contenedor

pudiendo recoger el humus líquido en la parte inferior del mismo. A diferencia de los

modulares, el vaciado del vermicompost debe realizarse mediante una puerta inferior o

lateral o, mediante volcado del contenedor con previa retirada de las lombrices mediante el

uso de trampas. A escala industrial se requieren un número importante de contenedores pero

en menor cantidad que las modulares, ya que se pueden emplear recipientes de mayor

volumen.

Introducción

44

Ambos sistemas permiten la recogida de fertilizante líquido y las lombrices están más

protegidas de las condiciones ambientales y de depredadores que el sistema de literas. Así

mismo, la superficie necesaria para un sistema industrial es menor que las literas (Tabla 10).

Tabla 10. Ventajas y desventajas del sistema de vermicompostaje en literas, modificado de

Edwards (2010).

Ventajas Mayor protección de las condiciones ambientales

Menor necesidad de espacio que las literas

Desventajas Gastos considerables en contenedores, recipientes y equipos

de movimiento de contenedores

Dificultad para mantener condiciones óptimas de humedad

Mano de obra intensiva

Dificultades para recoger el vermicompost sin lombrices en

no modulares

Lechos de flujo continuo

En general, son sistemas mecanizados que se emplean a escala media e industrial, y

requieren una inversión inicial mayor. Consisten en lechos o contenedores elevados por

encima del suelo con la base perforada y que presentan una descarga mecánica de las capas

inferiores de vermicompost (Figura 11).

Figura 11. Foto de un vermireactor diseñado por el Equipo de Biotecnología Ambiental de la

Universidad de Vigo (izquierda) y foto de una planta de producción de vermicompost de la

empresa Vermigrand (derecha).

Las capas de residuo se disponen periódicamente en la parte superior del lecho, de

manera que, las lombrices ascienden al alimento fresco y el vermicompost libre de lombrices

de tierra se acumula en la parte inferior. Esta separación se debe a que las lombrices epigeas

habitan la superficie, en torno a 10-15cm, y se mueven constantemente hacia la nueva fuente

de alimento, por lo que el sistema de flujo continuo elimina la necesidad de separar las

lombrices del vermicompost. La adicción de alimento puede ser mecanizada, así como, la

Introducción

45

recogida del vermicompost en la base del lecho. Al estar elevado del suelo está protegido de

depredadores y pueden mantenerse bajo cubierta para controlar las condiciones ambientales

(Tabla 11).

Tabla 11. Ventajas y desventajas del sistema de vermicompostaje en literas, modificado de

Edwards (2010).

Ventajas Alta protección de las condiciones ambientales y

depredadores

Menor necesidad de mano de obra

Mayor control de las condiciones de proceso

Posibilidad de automatización

Desventajas Importante inversión económica

OBJETIVOS

Objetivos

49

El objetivo general de este estudio es conocer la evolución de las actividades y las

comunidades microbianas en la fase de maduración del proceso de compostaje. Además, se

estudia esa misma dinámica en el proceso de vermicompostaje y se comparan en los dos

procesos de forma paralela y combinada. Las investigaciones se han enfocado en la fase de

maduración del compostaje con la finalidad de mejorar el proceso mediante el empleo de

distintas prácticas, entre ellas la inoculación de lombrices de tierra, y a la evaluación de la

estabilidad de los productos obtenidos a partir de los diferentes procesos. En consecuencia,

los objetivos específicos pueden resumirse como:

1. Obtener información que permita mejorar y optimizar el proceso de

compostaje a través del estudio de la estancia en maduración de material pre-compostado de

diferentes residuos y mediante distintas intervenciones (capítulos 1, 2, 3 y 4).

2. Estudiar el efecto que presenta el tipo de residuo sobre la evolución de la

dinámica microbiana, tanto de las comunidades microbianas como las actividades

enzimáticas implicadas en los procesos hidrolíticos de degradación de los compuestos

orgánicos, durante la estancia en maduración del proceso de compostaje (capítulo 1).

3. Establecer si la realización de prácticas de volteo sobre el material pre-

compostado en fase de maduración de un residuo energético es más efectivo para la

obtención de un producto es mejores condiciones de estabilidad y/o en menor tiempo que el

mantenimiento estático del compost (capítulo 2).

4. Ampliar el estudio sobre un residuo poco investigado como es el lodo de

depuradora proveniente de la elaboración de alimentos precocinados y ultracongelados del

mar (capítulos 1 y 2).

5. Determinar la viabilidad de la inoculación de lombrices de tierra epigeas en

residuos frescos y pre-compostados de diferente origen (capítulos 3 y 4).

6. Determinar cómo influye en la calidad de los productos y la evolución de los

procesos el empleo de lombrices de tierra tanto en el tratamiento de residuo orgánico fresco

como en la maduración del residuo pre-compostado a través del estudio de la dinámica

microbiana (capítulo 3 y 4).

7. Establecer si la maduración de estiércol pre-compostado mediante prácticas

dinámicas por medio de la inoculación de lombrices de tierra es más efectivo para la

obtención de un producto en mejores condiciones de estabilidad y/o en menor tiempo que

mediante maduración estática (capítulo 4).

Objetivos

50

8. Estudiar si las relaciones entre las actividades enzimáticas y las comunidades

microbianas permiten determinar hacia donde evolucionan los distintos tratamientos

estudiados y si permiten establecer las condiciones de estabilidad de los procesos orgánicos

(capítulos 1, 2, 3 y 4).

CAPÍTULO 1

Evolution of microbial dynamics during the maturation

phase of the composting of different types of waste

Villar I., Alves D., Garrido J., Mato S., 2016. Evolution of microbial dynamics during the

maturation phase of the composting of different types of waste. Waste Manag. 54, 83–

92. http://doi: 10.1016/j.wasman.2016.05.011.

Estado: publicado

Permiso: pdf editor

Capítulo 1

53

RESUMEN

Durante el compostaje, las instalaciones suelen ejercer un mayor control sobre la fase

bio-oxidativa del proceso, para la cual se utiliza una tecnología específica y, generalmente,

tiene una duración establecida. Tras esta fase, el material se deposita a madurar con menos

control. Si bien ha habido un estudio considerable de los parámetros biológicos durante la

fase termofílica, hay menos investigación sobre la fase de estabilización y maduración. Este

estudio evalúa los efectos del tipo de material de partida sobre la evolución de la dinámica

microbiana durante la fase de maduración del compostaje. Se utilizaron tres tipos de

residuos: lodo de la industria de transformación de pescado, lodo de aguas residuales

municipales y estiércol de cerdo, cada uno independientemente mezclado con madera de

pino triturada como agente estructurante. El sistema de compostaje, para cada residuo,

comprendió un reactor estático de 600L de capacidad para la fase bio-oxidativa seguida por

la fase de estabilización y maduración en triplicado en cajas de 200L durante 112 días. Los

ácidos grasos fosfolípidos, las actividades enzimáticas y los parámetros físico-químicos

fueron determinados a lo largo de la fase de maduración. La evolución de la biomasa

microbiana total, bacterias Gram +, bacterias Gram −, hongos y enzimas (β - glucosidasa,

celulasa, proteasa, fosfatasa ácida y alcalina) dependió significativamente del tipo de residuo

(p < 0.001). La comunidad microbiana predominante para cada tipo de residuo permaneció

durante todo el proceso de maduración, lo que indica que el tipo de residuo determina los

microorganismos que pueden desarrollarse en esta etapa. Mientras que los hongos

predominaron durante la maduración del lodo de pescado, el estiércol y el lodo municipal se

caracterizaron por una mayor proporción de bacterias. Tanto la estructura de la comunidad

microbiana como las actividades enzimáticas proporcionaron información importante para

monitorear el proceso de compostaje. Debe prestarse más atención a la fase de maduración

para optimizar el proceso de compostaje.

Palabras claves: actividad enzimática, ácidos grasos fosfolípidos (PLFA), comunidad

microbiana, compostaje, maduración

Waste Management 54 (2016) 83–92

Contents lists available at ScienceDirect

Waste Management

journal homepage: www.elsevier .com/locate /wasman

Evolution of microbial dynamics during the maturation phase of thecomposting of different types of waste

http://dx.doi.org/10.1016/j.wasman.2016.05.0110956-053X/� 2016 The Authors. Published by Elsevier Ltd.This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

⇑ Corresponding author.E-mail address: [email protected] (I. Villar).

Iria Villar ⇑, David Alves, Josefina Garrido, Salustiano MatoDepartment of Ecology and Animal Biology, University of Vigo, 36310 Vigo, Spain

a r t i c l e i n f o

Article history:Received 30 December 2015Revised 9 May 2016Accepted 10 May 2016Available online 25 May 2016

Keywords:Enzyme activityPhospholipid fatty acid (PLFA)Microbial communityCompostingMaturation

a b s t r a c t

During composting, facilities usually exert greater control over the bio-oxidative phase of the process,which uses a specific technology and generally has a fixed duration. After this phase, the material isdeposited to mature, with less monitoring during the maturation phase. While there has been consider-able study of biological parameters during the thermophilic phase, there is less research on the stabiliza-tion and maturation phase. This study evaluates the effects of the type of starting material on theevolution of microbial dynamics during the maturation phase of composting. Three waste types wereused: sludge from the fish processing industry, municipal sewage sludge and pig manure, each indepen-dently mixed with shredded pine wood as bulking agent. The composting system for each waste typecomprised a static reactor with capacity of 600 L for the bio-oxidative phase followed by stabilizationand maturation phase in triplicate 200 L boxes for 112 days. Phospholipid fatty acids, enzyme activitiesand physico-chemical parameters were measured throughout the maturation phase. The evolution of thetotal microbial biomass, Gram + bacteria, Gram � bacteria, fungi and enzymatic activities (b-glucosidase,cellulase, protease, acid and alkaline phosphatase) depended significantly on the waste type (p < 0.001).The predominant microbial community for each waste type remained present throughout the maturationprocess, indicating that the waste type determines the microorganisms that are able to develop at thisstage. While fungi predominated during fish sludge maturation, manure and municipal sludge were char-acterized by a greater proportion of bacteria. Both the structure of the microbial community and enzy-matic activities provided important information for monitoring the composting process. Moreattention should be paid to the maturation phase in order to optimize composting.� 2016 The Authors. Published by Elsevier Ltd. This is anopenaccess article under the CCBY-NC-ND license

(http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Composting is a process of biological degradation of solidorganic substrates under aerobic conditions through the action ofdifferent microbial populations, yielding a stable, humidified andsuitable product to add to the soil (Insam and de Bertoldi, 2007).The organic material goes through different phases: a mesophilicphase, characterized by the proliferation of the microbiota, a ther-mophilic phase where a high rate of biodegradation, the growth ofthermophilic organisms and the inhibition of non-thermotolerantorganisms occur and the final phase that includes a period of cool-ing, stabilization and maturation, characterized by the growth ofmesophilic organisms and the humification of the compost(Ryckeboer et al., 2003b). In the composting facilities the matura-tion phase is usually carried out with less control and monitoringthan the bio-oxidative phase. The duration of the bio-oxidative

phase that is carried out in bioreactors depends upon the type ofsubstrate that is used but generally lasts from 7 to 15 days. Afterthis phase, the material that exits the reactor generally is placedin windrows for a curing phase (Diaz et al., 2007). The timerequired for the maturation phase is a function of the substrateand environmental and operating conditions of the facility andcan range from a few weeks to a year or two (Diaz et al., 2002). Thislack of control over the process may cause environmental prob-lems such as odours and leachates, in addition to adversely affect-ing the quality of the compost.

The maturation phase has mainly been studied in terms of thephysico-chemical and biological parameters in order to determinewhen compost is mature enough to be added to the soil by estab-lishing maturity and stability criteria and indexes of the final pro-duct (Bernal et al., 2009; Insam and de Bertoldi, 2007; Paradeloet al., 2010). In terms of biological parameters, several enzymaticstudies have been carried out to determine microbiological activityduring composting and provide indicators of the stability of differ-ent composts (Cayuela et al., 2008; Ros et al., 2006). Castaldi et al.

Table 1Physico-chemical characteristics of the wastes used in the composting experiments:sludge from fish processing industry (FI), municipal sewage sludge (MSS) and pigmanure (PM).

FI MSS PM

Moisture (%) 65.4 87.0 82.8Organic matter (%) 93.0 73.1 79.6Total carbon (mg g�1 dw) 532.5 364.0 402.1Total nitrogen (mg g�1 dw) 26.8 46.5 31.9Water soluble carbon (mg g�1 dw) 20.50 2.54 19.78Dissolved organic nitrogen (mg g�1 dw) 10.05 9.07 10.18Total phosphorus (mg g�1 dw) 7.16 20.01 16.07Fats (% dw) 22.3 6.0 15.6

dw: dry weight.

84 I. Villar et al. /Waste Management 54 (2016) 83–92

(2008) proposed that the study of dynamics of certain enzymaticactivities, without single point determinations, could be a suitableindicator of stability, although it was not possible to establish athreshold value. However, the study of enzyme activities providesinformation on the breakdown of organic matter and the metabolicprocesses that take place during composting and, therefore, onproduct stability.

Most studies on the structure of the microbial community incomposting have specifically focused on the early stages of the pro-cess because, under aerobic conditions, temperature is the biggestselective factor of microbial populations (Ryckeboer et al., 2003b).Likewise, the nature of the organic substrates is also an importantfactor in determining the dynamics and microbial diversityduring composting (Klammer et al., 2008; Ryckeboer et al.,2003b; Vargas-García et al., 2010). Ishii and Takii (2003) showedthat the main factor affecting microbial communities was theconcentration of dissolved organic substances, which dependedon the type of starting material. López-González et al. (2015)showed that the composition of fresh materials and operatingconditions determine how the microbiota behaves, as well as itsstructure and its biodiversity.

Phospholipid fatty acid (PLFA) profiles analysis is a techniquethat provides information about the structure of the microbialcommunity and how it changes during composting. The totalamount of PLFAs can be used as an indicator of viable microbialbiomass. Furthermore, some PLFAs are specific to certain livingorganisms (e.g. bacteria, fungi, actinomycetes and plants) whichmeans they can be used as biomarkers for the presence and abun-dance of specific microbial groups (Zelles, 1999). There have beenstudies of the evolution of PLFAs during the composting of differ-ent wastes, with greatest emphasis on the initial stages of the pro-cess and some authors including on time sampling during thematuration phase (Amir et al., 2010; Eiland et al., 2001;Hellmann et al., 1997; Klamer and Bååth, 1998). Jindo et al.(2012) found that after 12 weeks of composting, factor analysisbased on the relative abundance of individual PLFAs revealedchanges in the structure of the microbial community thatdepended on the original organic waste. Boulter-Bitzer et al.(2006) studied the microbial community of different compostsduring maturation and storage, noting that PLFA analysis was avaluable method for characterizing the microbial communitystructure during this phase of the composting process.

The study of biological parameters during stabilization andmaturation and the influence of the source material can helpimprove the quality of compost and optimize the composting pro-cess, meaning that a better understanding of changes in the micro-bial dynamics is necessary for the maturation phase of composting.The objectives of this research were: (1) to study the developmentand structure of the microbial community using PLFAs and enzy-matic activities during the maturation phase of the compostingprocess; (2) evaluate the effect of the type of waste on the micro-bial structure and activity; and (3) improve and optimize the com-posting process by providing more information on the maturationphase.

2. Materials and methods

2.1. Composting materials

Composting experiments were performed using three differentwaste types whose main physico-chemical properties are detailedin Table 1:

– Sewage sludge from the food industry (FI), from precooked andfrozen fish and cephalopods, obtained after separation of fatsand treatment with coagulants and flocculants.

– Manure from a pig-breeding farm (PM), collecting the solidfraction of slurry after storage in manure pits.

– Municipal sewage sludge wastewater (MSS) obtained afteraerobic digestion in lagoons and dewatering with band filter.

Shredded pine wood passed through a 3 cm sieve was used as abulking agent and each waste type was mixed with this agent toobtain a ratio 1:2 (v/v), a free air space (FAS) of 30–40% and a mois-ture content of around 60–70%.

2.2. Experimental design

After applying the bulking agent, each waste type was subjectedto a composting process in which the bio-oxidative phasetook place in a static reactor with forced ventilation and thestabilization and the maturation phase (hereinafter referred to asthe maturation phase) were carried out in triplicate in 200 Lbatches (Fig. 1).

The adiabatic composting reactor had an effective volume of600 L, a perforated floor and a ventilation system with the abilityto introduce fresh or recirculated air through the top and bottomof the reactor. The temperature and oxygen level were recordedevery minute using a Eurotherm controller with three temperatureprobes at different depths and a gas probe inside the mass with anoxygen sensor. A feedback loop of oxygen and temperature (venti-lation when temperatures exceeded 60 �C or oxygen fell below 5%)and a time controller were used for aeration. The material was keptin the composting reactor until the temperature fell below 35 �C,requiring a total of 20, 18 and 17 days for FI, PM and MSS, respec-tively. After emptying the reactor, each waste type was mixed andplaced in triplicate in maturation systems of 200 L. Wooden boxesof 70 � 54 � 54 cm with a perforated base and open top were usedfor the maturation systems to allow gas exchange with the outside,attempting to simulate the inside of a maturation pile by maintain-ing some isolation from the outside. Similarly, to take placecorrectly, composting requires moisture balanced conditions toprevent the occurrence of water stress that might generate biolog-ical inactivity and false compost stabilization. A layer of bulkingagent was placed at the top and bottom of the box to provide ther-mal insulation for the waste and prevent moisture loss. Mesh wasplaced between the composted material and the bulking agent toprevent them mixing.

The beginning of the maturation process (day 0) started afteremptying the reactor. Maturation systems were emptied andmixed by hand at 14, 28, 42, 56, 70 and 91 days to simulate thedynamics of turning a pile of compost under maturation. A com-posite sample was taken of each system during each turning. Thetotal volume of the composite sample was 1500 mL. This wassieved through a 1 cm mesh to remove the bulking material. Thetemperature and oxygen level was monitored daily for the first42 days and three times per week until the end of the process at112 days. The moisture level was maintained above 60%, except

Centrifugal fan

Control andregistrationsystem

Temperature probesGas intake probe

plenum chamber

Ventilation system

Biofilter

bulking agent

compost

mesh

perforatedbase

woodenbox

Fig. 1. Scheme of the reactor and the maturation boxes used during the composting process for each waste type.

I. Villar et al. /Waste Management 54 (2016) 83–92 85

for FI, which required periodic rewetting in the first weeks of theprocess related to its highly hydrophobic nature due to a high lipidcontent (Table 1) and the higher temperatures reached at the startof the maturation phase.

2.3. Chemical analysis

The moisture and organic matter contents of the samples weredetermined after drying at 105 �C until constant weight and ashingat 550 �C for 4 h, respectively. Fresh samples were extracted with0.5 M K2SO4 in a ratio 1:10 (w/v) for analyses of inorganic nitrogen(N-NH4

+ and N-NO3�) (Sims et al., 1995) and dissolved organic nitro-

gen content (DON) (Cabrera and Beare, 1993). Total nitrogen con-tent (TN) and total carbon content (TC) were determined bycombustion of dried samples using a LECO 2000 CN elemental ana-lyzer. Water soluble carbon content (WSC) was analyzed in aque-ous extracts 1:5 (w/v) by dichromate oxidation in sulfuric acidsolution (Sims and Haby, 1971). Electrical conductivity was deter-mined in aqueous extracts 1:10 (w/v) using a conductivimeter Cri-son CM 35.

2.4. Biological and biochemical analysis

The microbial community composition and biomass weredetermined by phospholipid fatty acid (PLFA) analysis followingthe method described by Gómez-Brandón et al. (2010) for organicsamples. Briefly, total lipids were extracted by stirring from200 mg of each freeze-dried sample with 60 mL of chloroform–methanol (2:1, v/v). Phospholipid fraction was obtained after sep-aration on silicic acid columns and was subjected to derivatizationwith trimethylsulfonium hidroxyde (TMSH). Fatty acid methylesters (FAMEs) obtained were analyzed by gas chromatographyand mass spectrometry (GC–MS). GC–MS analysis was performedon a column CP-Select FAME, 100 m � 0.25 mm. FAMEs wereidentified by comparison of their retention time and mass spectrawith known standards (Larodan Fine Chemicals AB, Malmo,Sweden). The quantification was performed with methylnonadecanoate fatty acid (C19:0) as internal standard. PLFAs wereused to estimate the biomass of specific microbial groups:gram-positive bacteria (i14:0, i15:0, a15:0, i16:0, a17:0),gram-negative bacteria (16:1x7, 17:1x7, 18:1x7, cy19:0) andfungi (18:2x6, 18:1x9, 20:1x9). The total amount of PLFAsidentified (totPLFAs) was used as an indicator of the viable micro-bial biomass (Zelles, 1999).

b-glucosidase was estimated by incubating 1 g of fresh samplewith 1 mL of p-nitrophenyl-b-D-glucopiranoside (0.025 M) for 1 hat 37 �C and subsequent colorimetric measurement ofp-nitrophenol released (Eivazi and Tabatabai, 1988). Alkaline andacid phosphatase were measured by incubating 0.5 g of fresh sam-ple with 1 mL of p-nitrophenylphosphate (0.015 M) for 1 h at 37 �Cand subsequent colorimetric measurement of p-nitrophenolreleased (Eivazi and Tabatabai, 1977). Protease was measured bycolorimetric determination of the amino acids released, after theincubation of 1 g of fresh sample with 5 mL of sodium caseinate(2%) for 2 h at 50 �C, using Folin-Ciocalteu reagent (Ladd andButler, 1972). Cellulase was assessed by colorimetric determina-tion of reducing sugars released after incubation of 5 g of freshsample with 15 mL of carboxymethyl cellulose sodium salt (0.7%)for 24 h at 50 �C (Schinner and von Mersi, 1990).

Germination index (GI) was calculated according to Zucconiet al. (1981) by determining seed germination and root length ofLepidium sativum growing in 2 mL of aqueous extracts 1:5 (w/v)in Petri dishes lined with paper filter during 48 h.

Static respiration rate (SR) was measured using manometricrespirometers by OxiTop� system (WTW GmbH, Weilheim,Germany). Briefly, fresh weight equivalent to 4 g of dried samplewas placed in a hermetic container with a 1 M NaOH trap tocapture CO2, and the pressure drop, due to microbial oxygenconsumption, was recorded during 24 h at constant temperature.The self-heating test was carried out using 2 L Dewar flask for10 days at room temperature (TMECC, 2002).

2.5. Statistical analysis

All statistical tests were performed using R software (RDevelopment Core Team, 2014). The physico-chemical data wassubjected to principal component analysis (PCA) after normaliza-tion to zero mean and unit variance, and the analysis wasperformed on the correlation matrix. The principal componentswith eigenvalues greater than one were retained. The PCA wasperformed using the prcomp function and the package ggbiplot(Vu, 2011). Enzyme and PLFA data was analyzed with linearmixed-effects models using the nlme package (Pinheiro et al.,2015). The waste type and time were fixed factors and the repeatedmeasurement throughout time in each maturation box was treatedas a random effect to address the non-independence of samples.Logarithmic and square root transformations of the data were nec-essary to ensure the normality and homogeneity of the variance of

86 I. Villar et al. /Waste Management 54 (2016) 83–92

residuals of models. For post hoc comparison between treatments,Tukey tests were carried out using the glht function of the mult-comp package (Hothorn et al., 2008). Correlation analyses werealso carried out to examine the relationships between PLFAs andenzyme activities with cor function of the stats package. All statis-tical tests were evaluated at the 95% confidence level and valuesare given as the mean ± standard error.

3. Results

3.1. Temperature evolution

The temperature inside the reactor increased rapidly at thebeginning of the process (Fig. 2), reaching over 45 �C on days oneand two for FI and PM, respectively, and both wastes maintainedthermophilic temperatures for more than 10 days. For MSS, ther-mophilic temperatures were reached inside the reactor on day fourand were maintained for seven days.

At the beginning of the maturation phase, the temperature of FIincreased to 60 �C, falling on day four and reactivating after turningat 14 and 28 days, although in the latter cases, the temperatureremained below 40 �C. Both PM and MSS generally remainedbelow 25 �C during the maturation phase. In terms of the oxygenlevels during the maturation phase, measurements remainedabove 17% for all waste types.

3.2. Composting parameters

The principal component analysis of the physico-chemical vari-ables is presented in Fig. 3 and shows the three separate groups forthe three different waste types. The samples taken during matura-tion of FI exhibited the greatest dispersion along the axes. Theparameters responsible for the differences between waste typesalong principal component one (PC1) were WSC (r = �0.97,p < 0.0001) and moisture (r = 0.89, p < 0.0001), differentiating sam-ples during the maturation of FI, PM and MSS. During the process,FI maintained a high concentration of WSC, with a maximum of25.6 mg g�1 at the beginning of the maturation phase and a mini-mum of 9.1 mg g�1 at 112 days. Both PM and MSS reached a max-imum of around 7.5 mg g�1, with a drop throughout thematuration process to achieve significantly lower values in PMthan MSS (Table 2). In terms of moisture, PM and MSS remainedat around 60–70% throughout the process, while FI required peri-odic rewetting during the first seven weeks, obtaining minimumvalues of 35% by day 28 and remaining between 45% and 55% fromday 42 until the end of the process. The parameter with the biggestcontribution to the separation of the waste along principal compo-nent two (PC2) was the ratio C/N (r = 0.92, p < 0.0001), primarily

Fig. 2. Temperature evolution during the reactor phase and the maturation phasefor sludge from fish processing industry (FI), municipal sewage sludge (MSS) andpig manure (PM). Arrows correspond to the turnings after emptying the reactor(time 0) and during the maturation phase (14, 28, 42, 56, 70, 91 days).

differentiating between PM and MSS. In the latter one, values ofC/N > 12 were not detected along the maturation phase, whilePM and FI reached around 18 at the beginning of maturation, withsignificantly lower values for FI at the end of the process (Table 2).

3.3. Microbial community

Microbial biomass, Gram + bacteria, Gram � bacteria and fungiwere significantly different between wastes (p < 0.001). Likewise,significant differences caused by time (p < 0.0001) and significantinteractions between time and waste type (p < 0.0001) for allmicrobial biomass and microbial groups were observed.

The highest concentration of microbial biomass was observed inMSS at the beginning of maturation and fell progressively overtime (Fig. 4). This reduction was also present in the group indica-tors, falling to about 97.8% in Gram + bacteria, 99.5% in Gram �bacteria and 99.3% in fungi. At the beginning of the process theproportion of PLFAs of a specific group with respect to total PLFAsgroup indicators was higher in Gram� bacteria (>40%), whereas bythe end, the greater proportion was in Gram + bacteria (>60%).Hence, more PLFAs characteristic of bacteria than fungi ones wereobserved during the process. Finally, the microbial biomass of MSSduring maturation was correlated with all microbial groups(r > 0.98, p = 0.000).

In the case of FI, there was an increase in microbial biomassduring the first weeks of the process, declining in the followingsamplings and showing a significant increase during the last21 days of maturation. The microbial biomass was strongly corre-lated with the concentration of fungi (r = 0.915, p = 0.000). In thiscase, the predominance of PLFAs characteristic of fungi was main-tained at over 55% during the process.

Furthermore, PM presented less microbial biomass than theother ones in the first samplings, gradually falling until day 56and later recovering to values slightly below the initial value. Ahigh correlation between microbial biomass and the PLFAs charac-teristic of Gram + and Gram � bacteria (r > 0.93, p < 0.0001) wereobserved. During the maturation phase, bacteria predominated(values above 75%), especially Gram + bacteria.

3.4. Enzyme activities

The type of waste significantly affected the evolution of allhydrolytic activities studied (p < 0.001). Also, significant differ-ences caused by time (p < 0.0001) and interaction between timeand waste (p < 0.0001) were also observed for all the enzymes.

After intensive composting in the reactor, MSS presented thehighest activity of b-glucosidase (Fig. 5a), which fell significantlyin the first 14 days and remained stable from 70 days, resultingin a final reduction of 85%. In contrast, at the start of the matura-tion phase, PM and FI showed similar values of b-glucosidase,but from day 56, PM had significantly higher values than the otherwaste types.

With respect to cellulase (Fig. 5b), increased activity wasobserved in PM, except for the last sampling where similar resultswere found in FI. In the case of MSS, a continued reduction in cel-lulase activity was detected, while it increased slightly for FI in thefinal maturation sampling.

The acid and alkaline phosphatase activities exhibited similartrends for the same waste type (Fig. 5c and d, respectively), exceptthe initial samplings of PM. Acid phosphatase activity was similarto b-glucosidase activity, with higher levels in MSS for initial sam-plings and high activity at the end of the process for PM. However,FI remained significantly lower throughout maturation. Despitethe different activity levels observed at the reactor outlet, FI andMSS showed similar trends of alkaline phosphatase enzymes, withactivity peaking on day 42, followed by a sharp decline to similar

Fig. 3. Correlation biplot of principal component analysis where vectors correspond to the variables that define the components and points correspond to sampling duringthe maturation phase for sludge from fish processing industry (FI), municipal sewage sludge (MSS) and pig manure (PM).

Table 2Parameters of composts after 112 days of maturation phase for sludge from fishprocessing industry (FI), municipal sewage sludge (MSS) and pig manure (PM).

FI MSS PM

Ratio C/N 12.25 ± 0.11a 11.53 ± 0.20b 16.31 ± 0.31c

Ratio NH4+/NO3

� 0.24 ± 0.03a 0.09 ± 0.01b 0.56 ± 0.03c

SR (mg O2 g�1 OM h�1) 0.53 ± 0.01a 0.33 ± 0.02b 0.78 ± 0.01c

Self-heating test Class V Class V Class VGI (%) 91.3 ± 2.1a 99.5 ± 1.1b 94.1 ± 1.4a

WSC (mg g�1 dw) 9.11 ± 0.15a 3.50 ± 0.20b 2.67 ± 0.15c

SR: static respiration, GI: germination index, WSC: water soluble carbon, dw: dryweight.In each parameter the different letters indicate significant differences betweencomposts (Tukey post hoc test p < 0.05).

I. Villar et al. /Waste Management 54 (2016) 83–92 87

values for both waste types at the end of the process. Similar toacid phosphatase and b-glucosidase, PM showed significantlyhigher values of alkaline phosphatase than MSS and FI in the finalsamplings.

Only the protease enzyme (Fig. 5e) exhibited similar trends inall wastes, with reduction rates of 80%, 72% and 22% for PM, MSSand FI respectively. In all cases, activity stabilized in the finalsamplings.

As shown in Table 3, microbial groups for MSS were positivelycorrelated with all enzymatic activities. In contrast, in FI, onlyPLFAs characteristic of fungi were positively correlated with theb-glucosidase and alkaline phosphatase enzymes. With respect tobacterial biomass for PM, both Gram + and Gram � bacteria werecorrelated with alkaline phosphatase, protease and cellulase activ-ities, while fungal biomass was positively correlated with cellulase.

4. Discussion

4.1. Temperature evolution

Unlike FI, PM and MSS exhibited biological degradation oforganic matter prior to composting, the former during storage inseptic tanks for solid-liquid separation of the pig slurry and the lat-

ter during treatment of wastewater in aeration tanks. Conse-quently, both had a lower content of organic matter than FI(Table 1), allowing the full thermophilic phase to take place inthe static reactor. The materials did not require turning to reacti-vate the process, showing that the forced ventilation was effective,and both fresh composts matured in boxes with stable and envi-ronmental temperatures. In the case of FI, however, turningfavoured the increase in temperature at the beginning (day 0),and at 14 and 28 days, meaning that forced ventilation was not suf-ficient to allow full development of the bio-oxidative phase of thecomposting process in the reactor. Alburquerque et al. (2009)showed that forced ventilation was effective when performedtogether with mechanical turning during alperujo (olive waste)composting, concluding that turning improved the porosity andhelped distribute the moisture, substrates and microorganisms.During treatment of wastewater from fish processing a digestionprocess was not performed, meaning that the organic load of thesludge was high (Table 1), and turning and watering during thefirst weeks allowed re-heating to provide biodegradable substratesfor the microorganisms from the outside to the inside of the box.Despite this re-heating, FI kept the characteristics of a maturationprocess with temperatures similar to MSS and PM from the firstmonth. So, in a composting facility, this highly organic wasteshould be monitored when it is disposed to mature, to preventfalse stabilizations of compost, odours and other environmentalproblems.

4.2. Multivariate analysis

Multivariate analysis shows that the source material deter-mined the physico-chemical development of the stabilization andmaturation process for composting. Several authors have foundthat composts obtained from different organic waste types differin their physico-chemical composition and, hence, in terms of theirstabilities and qualities, depending on the composition of sourcematerial used in the composting (Bernal et al., 1998; Ranalliet al., 2001). Similarly, the physico-chemical properties during

Fig. 4. Changes in (a) microbial biomass, (b) fungi, (c) Gram + bacteria and (d) Gram � bacteria estimated by phospholipid fatty acid (PLFA) analysis during the maturationphase for sludge from fish processing industry (FI), municipal sewage sludge (MSS) and pig manure (PM).

88 I. Villar et al. /Waste Management 54 (2016) 83–92

the maturation phase determined the separation of the three wastetypes and affected the final composition of the composts. FIshowed the greatest changes in physico-chemical compositionduring maturation, possibly due to its initial re-activation causedby its high organic load, as observed in the temperature profile.The highest respiratory activities and contents of C and N formsoccurred in FI, meaning more time may be required for stabiliza-tion. The maturation process of PMwas characterized by maintain-ing the C/N ratio at a low value (Bernal et al., 2009), suggesting apossible stabilization of the degradation process of organic matter.However, Fialho et al. (2010) have shown that the C/N ratio is not agood method for monitoring the composting process and thatthere is not an optimal ratio that characterizes humified compost.Finally, MSS was characterized by high levels of ammonia due tothe low C/N ratio present throughout the whole process. DeGuardia et al. (2008) have shown that an excess of aeration afterthe thermophilic phase could be responsible for the loss of N. Thus,stabilization of the composting parameters of MSS suggested thatthere were too much turnings or time processing for this type ofwaste, with the potential to increase the volatility and loss of Nduring the maturation phase.

4.3. Compost quality

All compost types presented parameters of stability and matu-rity as were highest rating during the self-heating test (class V,mature compost) and values of GI above 80% indicative of absenceof phytotoxic substances for plant growth (Riffaldi et al., 1986).However, MSS showed a higher degree of maturity and stabilitybecause it had a respiratory rate below 0.5 mg O2 g�1 SV h�1

(Iannotti et al., 1993), a ratio C/N below 12 and an ammonia/nitrateratio below 0.16 (Bernal et al., 1998). Likewise, Garcia et al. (1991)proposed the use of WSC content as a useful method for determin-ing the maturity of a compost, with a value below 5 mg g�1 formature compost. In this experiment, FI exhibited higher values ofthis content. Both PM and FI may require more time to achieve

similar maturity and stability parameters to MSS, even thoughboth were classified as ‘‘mature” according to the rates in TMECC(TMECC, 2002). Although the waste types studied underwent thesame treatment in composting, they exhibited different stabilityand maturity levels. As such, it is important to design and controlthe composting process as a whole, paying particular attention tothe maturation phase, where it is possible to design an ad hoc pro-cess to yield higher quality compost in less time, depending on thephysico-chemical properties of each type and its evolution overtime.

4.4. Microbiological evolution

Changes in both enzyme activities and the microbial commu-nity during the maturation phase were determined by the individ-ual features of organic waste types.

The decline in microbial biomass in MSS was consistent withprevious studies (Garcia et al., 1992; Klamer and Bååth, 1998;Mondini et al., 2004; Ros et al., 2006) using different methods ofquantification (Biolog, ATP, fumigation-extraction, PLFA) andmicrobial monitoring during the maturation phase. As mentioned,previous research has focused on the most intensive phase of com-posting, although some studies include occasional samplings dur-ing the maturation phase. Garcia et al. (1992) observed that thecontinued decrease in microbial biomass during the compostingof municipal solid waste could be attributed to the slow andincomplete stabilization of organic matter. Here it should be notedthat the low C/N ratio observed throughout the maturation of MSScould cause a shortage of available carbon, leading to a progressivedecline in microbial biomass, slowing down the process for thedegradation of organic matter. However, the results in Table 1show optimal properties for compost, suggesting the drop inmicrobial biomass was the result of the maturity and stability ofthe organic matter. In terms of the structure of the microbial com-munity, the abundance of PLFAs characteristic of bacteria detectedthroughout the maturation phase was high and could be attributed

Fig. 5. Changes in (a) b-glucosidase, (b) cellulase, (c) acid phosphatase, (d) alkaline phosphatase and (e) protease during the maturation phase for sludge from fish processingindustry (FI), municipal sewage sludge (MSS) and pig manure (PM).

I. Villar et al. /Waste Management 54 (2016) 83–92 89

to the origin and physico-chemical properties of the waste. Themain and most important microbial population that develops inwastewater treatment systems is bacteria, especially Gram � bac-teria such as Proteobacteria and Bacteroidetes, and Gram + bacteriasuch as Actinobacteria (Wagner and Loy, 2002). Furthermore, fol-lowing the phase of thermophilic composting, the recolonizationof the compost with mesophilic microorganisms of the environ-ment occurs, dominated by the bacteria of the phylum Bacteroide-tes (Insam and de Bertoldi, 2007; Ryckeboer et al., 2003b). Eilandet al. (2001) have found that bacteria dominated the microbialcommunity throughout the composting process when C/N waslow, although all treatments studied showed a fungi/bacteria ratioof less than 0.5. The waste used (straw with various additions ofslurry) was a predominantly bacterial medium for all tested mix-tures and throughout composting. Ishii and Takii (2003) observedsimilar bacterial communities in different composting sewagesludge processes, including Bacillus, Actinobacteria and Gram �bacteria. These authors have suggested that microorganisms thatproliferate in composting processes adapt to the composting envi-ronment and are selected by factors within the composting mate-rials. Furthermore, in a previous study using sewage sludge fromthe same wastewater plant, a predominance of PLFAs typical of

bacteria were observed in both vermicomposting and the com-bined process composting-vermicomposting (Villar et al., 2016).So, the physico-chemical factors exhibited by MSS characterizedthe microbial biomass that developed during the maturationprocess.

The microbial biomass for FI developed in a similar way to pre-vious studies on PLFAs (Boulter-Bitzer et al., 2006; Hellmann et al.,1997), with an increase in the abundance of PLFAs in the finalstages of composting. The microbial biomass growth in FI wascaused by an increase in PLFA fungi biomarkers. It is normal forfungi to increase during the maturation phase of composting as aresult of the breakdown of substrates of difficult degradation andless aggressive environmental factors (Albrecht et al., 2010;Hassen et al., 2001) and there is evidence of an increase in fungaldiversity (Shemekite et al., 2014). However, FI showed greaterabundance of PLFA fungi biomarkers than bacterial ones through-out the maturation process. The lipidic nature of the starting mate-rial and its highWSC content throughout maturation might involvethe proliferation of saprophytic sugar fungi, such as Zygomycetesspecies, in early stages of the maturation and the proliferation ofcellulolytic fungi during the final phase (Richardson, 2009;Ryckeboer et al., 2003b). Similarly, Amir et al. (2010) found that

Table 3Correlation matrix between enzyme activities and PLFAs during the maturation phasefor sludge from fish processing industry (FI), municipal sewage sludge (MSS) and pigmanure (PM).

totPLFAs Gram + Gram � Fungi

F1b-Glucosidase 0.358* NS �0.531** 0.384*

Cellulase NS �0.411* NS NSAcid phosphatase NS NS NS NSAlkaline phosphatase NS NS �0.421* 0.409*

Protease NS NS �0.627** NSMSSb-Glucosidase 0.921** 0.887** 0.900** 0.911**

Cellulase 0.892** 0.850** 0.889** 0.888**

Acid phosphatase 0.549** 0.582** 0.464** 0.564**

Alkaline phosphatase 0.702** 0.704** 0.634** 0.719**

Protease 0.821** 0.827** 0.756** 0.831**

PMb-Glucosidase NS NS NS NSCellulase 0.581** 0.459* 0.506** 0.590**

Acid phosphatase NS NS NS �0.401*

Alkaline phosphatase 0.644** 0.715** 0.633** NSProtease 0.720** 0.702** 0.601** NS

NS: not significant.* Indicates significance at the 0.05 probability level.

** Indicates significance at the 0.01 probability level.

90 I. Villar et al. /Waste Management 54 (2016) 83–92

the presence of fungi was greater in waste with a high fatty acidcontent, which impeded the stabilization and maturation of thecompost. Hence, the maintenance of high content of fungi couldbe indicative of the lack of stabilization during the maturationphase of this type of waste.

For PM, the stabilization of microbial biomass at the end of mat-uration suggested nutrients were available to the microorganisms,probably due to the high content of cellulosic materials, such asstraw and remnants of seeds that are normally present in pig slurryand degrade more slowly. Tiquia et al. (2002a) found ATP stabiliza-tion during pruning composting, indicating the maturity of thecompost and suggesting a change in the microbial community tomore specialized microbial groups, such as fungi and Actino-mycetes. However, as observed for MSS, bacterial biomass wasgreater than fungal biomass during the maturation process, partic-ularly PLFA biomarkers of Gram + bacteria. Elouaqoudi et al. (2015)have suggested that the increase in Gram + bacteria in the finalphase of composting indicates the availability of organic substratesdue to the breakdown of lignocellulosic compounds. Moreover, theorigin of waste, which characterizes its physico-chemical proper-ties, may influence the development of the microbial communityduring the maturation phase, since pig manure is an environmentrich in bacteria with a high content of fermentative microbialgroups, especially Gram + bacteria, and bacteria dominate duringthe initial phase of the decomposition of manure due to the highavailability of water and compounds that can be easily brokendown (Domínguez et al., 2010; Snell-Castro et al., 2005).

The change in the microbial community during the maturationprocess appears to be affected by the adaptation of microorgan-isms to mesophilic conditions and the different physico-chemicalproperties of different sources of waste, meaning there is a contin-uous turnover of the microbial groups while maintaining the influ-ence of the nature of the starting material and the predominantmicrobial groups.

The study of enzymatic activities during maturation showed thedynamics of the metabolic processes of the carbon, nitrogen andphosphorus cycle and thus hydrolytic enzymes were indicative ofthe evolution of organic matter and the biological activity duringthe final phase of the composting process.

In terms of MSS, the high correlation between all enzymes andmicrobial biomass indicates the enzyme activities during

maturation were a direct result of the microbial community, bothbacterial and fungal. The sharp decline of enzymes and PLFA con-tent over time suggests a decrease of available substrates formicroorganisms. At the beginning of the process, MSS had thehighest b-glucosidase and phosphatase activities. The low C/N ofthis waste may require the synthesis of b-glucosidase, since thelimitation of carbon with respect to the available nitrogen andphosphorus provides a strong incentive for microorganisms toinvest in the acquisition of carbon (Allison and Vitousek, 2005).Likewise, García et al. (1993) indicated that high values ofphosphatase activity in sewage sludge can be induced by thepresence of phosphate compounds from detergents in wastewater.b-glucosidase, protease and acid phosphatase enzymes showedstabilization in the final samplings while alkaline phosphatasecontinued to decline and cellulase increased slightly in the finalsampling. Castaldi et al. (2008) observed a decrease in bothenzyme activities and water-soluble fractions during compostingof the organic fraction of municipal solid waste with plant waste,indicating a stabilization of hydrolytic enzymes and hence theorganic matter during the maturation phase.

In contrast to MSS, no correlation was observed between themeasure of microbial biomass and enzymes, except forb-glucosidase during maturation of FI, suggesting the enzymeactivities were not directly associated with the development ofthe microbial biomass. Burns (1982) proposed that the enzymesmay be located in dead cells, cell debris or stabilized in organiccomplexes and remain as active hydrolyzing substrates. A furtherstudy by Mondini et al. (2004) found no correlation between themicrobial biomass carbon and different enzyme activities duringthe composting of different wastes, indicating stabilization ofextracellular enzymes due to the formation of complexes withhumic substances. During maturation, FI exhibited low enzymeactivities, compared to MSS and PM. The high content of nutrientsdirectly available for microorganisms, such as WSC and DON, couldinhibit the production of enzymes. Cellulase and protease activitiesincreased at the end of the process, indicating that the fall insoluble organic matter is accompanied by an increase in thehydrolysis of more complex organic compounds (Tiquia et al.,2002b).

In terms of the hydrolytic activity of PM, particularly high cellu-lase activity was observed throughout the maturation. However,this was in line with expectations of high cellulosic activity duringcomposting of pig manure due to the predominance of celluloseand hemicellulose in this type of waste (Iannotti et al., 1979). Bothbacteria and fungi determined cellulase activity and the predomi-nance of bacteria throughout the process resulted in the presenceof a significant cellulolytic bacterial community during maturationof PM, as was observed by Ryckeboer et al. (2003a) during the com-posting of vegetable, fruit and garden waste. Moreover, b-glucosidase and phosphatases increased at the end of the matura-tion process, with a notable progressive increase in alkaline phos-phatase over time. The highest level of enzymatic activity in thePM compost was similar to that shown by other authors(Cayuela et al., 2008; Tiquia and Tam, 1998) and was indicativeof the presence of available nutrients for the microbial biomass,especially bacteria, and/or the protection of enzymes in humiccomplexes, as indicated by the lack of correlation between b-glucosidase and acid phosphatase with the microbial community.The protease enzyme was of particular importance, since itdecreased during the maturation of PM, FI and MSS and stabilizedat the end of the process. This was the only enzyme to be positivelycorrelated with WSC (p < 0.01) in the three waste types, suggestingthe availability of carbon regulates the synthesis of proteases(Allison and Vitousek, 2005). Indeed, Lazcano et al. (2008) sug-gested that protease may be indicative of the degradation oforganic matter because of an extreme dependence on the availabil-

I. Villar et al. /Waste Management 54 (2016) 83–92 91

ity of substrate. Hence, the increased availability of substrates andtherefore the higher protease activity at the end of the process wasobserved in FI, PM and finally MSS.

The different correlation between the microbial groups andenzymatic activities during maturation provided informationon the state of the degradation of organic matter andcompost quality, as well as the requirement for more or lessprocessing time and handling. The authors recommend moreresearch on different maturation systems and waste typeswith particular attention to biological parameters during thematuration phase.

5. Conclusions

Taken together, the structure of the microbial community andenzymatic activities provide important information for monitor-ing the composting process and on the stability and maturity ofcompost. Low enzymatic activity is not indicative of stabilization,as observed by the high microbial community present in sludgefrom the fish processing industry, although a decrease in bothenzymatic activity and microbial community may indicate stabil-ity, as was the case during the maturation of municipal sewagesludge. The predominant microbial community for each type ofwaste remained present during the maturation process, indicatingthat the different origin of the waste and the different physico-chemical properties determine the microorganisms that are ableto develop at this stage. Although waste types were subjectedto the same composting process, the level of stability andmaturity was different and this represents an important factorfor designing ad hoc processes and controlling the compostingprocess according to waste type, paying particular attentionto the process as a whole, including maturation. It is importantto monitor microbial communities and their activity overtime to determine if and when compost is stable enough to beapplied to soil, or whether more time or alternative processmanagement is required.

Acknowledgments

This study was financially supported by the Xunta de Galicia(Regional Autonomous Government of Galicia) (09MDS024310PR).The authors thank the research support services of the Universityof Vigo (CACTI) for the carbon and nitrogen analysis. The authorsalso thank Emilio Rodríguez Cochón and Domingo Pérez Díaz fortechnical support.

References

Albrecht, R., Périssol, C., Ruaudel, F., Petit, J.L., Terrom, G., 2010. Functional changesin culturable microbial communities during a co-composting process: carbonsource utilization and co-metabolism. Waste Manage. 30, 764–770. http://dx.doi.org/10.1016/j.wasman.2009.12.008.

Alburquerque, J.A., Gonzálvez, J., Tortosa, G., Baddi, G.A., Cegarra, J., 2009. Evaluationof ‘‘alperujo” composting based on organic matter degradation, humificationand compost quality. Biodegradation 20, 257–270. http://dx.doi.org/10.1007/s10532-008-9218-y.

Allison, S.D., Vitousek, P.M., 2005. Responses of extracellular enzymes to simple andcomplex nutrient inputs. Soil Biol. Biochem. 37, 937–944. http://dx.doi.org/10.1016/j.soilbio.2004.09.014.

Amir, S., Abouelwafa, R., Meddich, A., Souabi, S., Winterton, P., Merlina, G., Revel, J.-C., Pinelli, E., Hafidi, M., 2010. PLFAs of the microbial communities incomposting mixtures of agro-industry sludge with different proportions ofhousehold waste. Int. Biodeterior. Biodegrad. 64, 614–621. http://dx.doi.org/10.1016/j.ibiod.2010.01.012.

Bernal, M.P., Alburquerque, J.A., Moral, R., 2009. Composting of animal manures andchemical criteria for compost maturity assessment. A review. Bioresour.Technol. 100, 5444–5453. http://dx.doi.org/10.1016/j.biortech.2008.11.027.

Bernal, M.P., Paredes, C., Sánchez-Monedero, M.A., Cegarra, J., 1998. Maturity andstability parameters of composts prepared with a wide range of organic wastes.Bioresour. Technol. 63, 91–99. http://dx.doi.org/10.1016/S0960-8524(97)00084-9.

Boulter-Bitzer, J.I., Trevors, J.T., Boland, G.J., 2006. A polyphasic approach forassessing maturity and stability in compost intended for suppression of plantpathogens. Appl. Soil Ecol. 34, 65–81. http://dx.doi.org/10.1016/j.apsoil.2005.12.007.

Burns, R.G., 1982. Enzyme activity in soil: location and a possible role in microbialecology. Soil Biol. Biochem. 14, 423–427. http://dx.doi.org/10.1016/0038-0717(82)90099-2.

Cabrera, M.L., Beare, M.H., 1993. Alkaline persulfate oxidation for determining totalnitrogen in microbial biomass extracts. Soil Sci. Soc. Am. J. 57, 1007–1012.http://dx.doi.org/10.2136/sssaj1993.03615995005700040021x.

Castaldi, P., Garau, G., Melis, P., 2008. Maturity assessment of compost frommunicipal solid waste through the study of enzyme activities and water-solublefractions. Waste Manage. 28, 534–540. http://dx.doi.org/10.1016/j.wasman.2007.02.002.

Cayuela, M.L., Mondini, C., Sánchez-Monedero, M.A., Roig, A., 2008. Chemicalproperties and hydrolytic enzyme activities for the characterisation of two-phase olive mill wastes composting. Bioresour. Technol. 99, 4255–4262. http://dx.doi.org/10.1016/j.biortech.2007.08.057.

de Guardia, A., Petiot, C., Rogeau, D., Druilhe, C., 2008. Influence of aeration rate onnitrogen during blackwater composting. Waste Manage. 248, 575–587. http://dx.doi.org/10.1016/j.wasman.2007.02.007.

Diaz, L.F., Savage, G.M., Eggerth, L.L., Chiumenti, A., 2007. Systems used incomposting. Compost Sci. Technol. 8, 67–87.

Diaz, L.F., Savage, G.M., Golueke, C.G., 2002. Composting of municipal solid wastes.In: Tchobanoglous, G., Kreith, F. (Eds.), Handbook of Solid Waste Management.McGraw-Hill Inc., New York, pp. 12.1–12.70.

Domínguez, J., Aira, M., Gómez-Brandón, M., 2010. Vermicomposting: earthwormsenhance the work of microbes. In: Insam, H., Franke-Whittle, I.H., Goberna, M.(Eds.), Microbes at Work: From Wastes to Resources. Springer, Heidelberg, pp.93–114.

Eiland, F., Klamer, M., Lind, A.-M., Leth, M., Bååth, E., 2001. Influence of initial C/Nratio on chemical and microbial composition during long term composting ofstraw. Microb. Ecol. 41, 272–280. http://dx.doi.org/10.1007/s002480000071.

Eivazi, F., Tabatabai, M.A., 1988. Glucosidases and galactosidases in soils. Soil Biol.Biochem. 20, 601–606. http://dx.doi.org/10.1016/0038-0717(88)90141-1.

Eivazi, F., Tabatabai, M.A., 1977. Phosphatases in soils. Soil Biol. Biochem. 9, 167–172. http://dx.doi.org/10.1016/0038-0717(77)90070-0.

Elouaqoudi, F.Z., El Fels, L., Amir, S., Merlina, G., Meddich, A., Lemee, L., Ambles, A.,Hafidi, M., 2015. Lipid signature of the microbial community structure duringcomposting of date palm waste alone or mixed with couch grass clippings. Int.Biodeterior. Biodegrad. 97, 75–84. http://dx.doi.org/10.1016/j.ibiod.2014.08.016.

Fialho, L.L., Lopes da Silva, W.T., Milori, D.M.B.P., Simões, M.L., Martin-Neto, L., 2010.Characterization of organic matter from composting of different residues byphysicochemical and spectroscopic methods. Bioresour. Technol. 101, 1927–1934. http://dx.doi.org/10.1016/j.biortech.2009.10.039.

García, C., Hernández, T., Costa, C., Ceccanti, B., Masciandaro, G., Ciardi, C., 1993. Astudy of biochemical parameters of composted and fresh municipal wastes.Bioresour. Technol. 44, 17–23. http://dx.doi.org/10.1016/0960-8524(93)90202-M.

Garcia, C., Hernández, T., Costa, F., 1991. Changes in carbon fractions duringcomposting and maturation of organic wastes. Environ. Manage. 15, 433–439.http://dx.doi.org/10.1007/BF02393889.

Garcia, C., Hernández, T., Costa, F., Ceccanti, B., Ciardi, C., 1992. Changes in ATPcontent, enzyme activity and inorganic nitrogen species during composting oforganic wastes. Can. J. Soil Sci. 72, 243–253. http://dx.doi.org/10.4141/cjss92-023.

Gómez-Brandón, M., Lores, M., Domínguez, J., 2010. A new combination ofextraction and derivatization methods that reduces the complexity andpreparation time in determining phospholipid fatty acids in solidenvironmental samples. Bioresour. Technol. 101, 1348–1354. http://dx.doi.org/10.1016/j.biortech.2009.09.047.

Hassen, A., Belguith, K., Jedidi, N., Cherif, A., Cherif, M., Boudabous, A., 2001.Microbial characterization during composting of municipal solid waste.Bioresour. Technol. 80, 217–225. http://dx.doi.org/10.1016/S0960-8524(01)00065-7.

Hellmann, B., Zelles, L., Palojärvi, A., Bai, Q., 1997. Emission of climate-relevant tracegases and succession of microbial communities during open-windrowcomposting. Appl. Environ. Microbiol. 63, 1011–1018.

Hothorn, T., Bretz, F., Westfall, P., 2008. Simultaneous inference in generalparametric models. Biom. J. 50, 346–363. http://dx.doi.org/10.1002/bimj.200810425.

Iannotti, D.A., Pang, T., Toth, B.L., Elwell, D.L., Keener, H.M., Hoitink, H.A.J., 1993.A quantitative respirometric method for monitoring compost stability.Compost Sci. Util. 1, 52–65. http://dx.doi.org/10.1080/1065657X.1993.10757890.

Iannotti, E.L., Porter, J.H., Fischer, J.R., Sievers, D.M., 1979. Changes in swine manureduring anaerobic digestion. Dev. Ind. Microbiol. 20, 519–529.

Insam, H., de Bertoldi, M., 2007. Microbiology of the composting process. In: Diaz, L.F., de Bertoldi, M., Bidlingmaier, W., Stentinford, E. (Eds.), Compost Science andTechnology. Waste Management Series. Elsevier Ltd., pp. 25–48, http://dx.doi.org/10.1016/S1478-7482(07)80006-6.

Ishii, K., Takii, S., 2003. Comparison of microbial communities in four differentcomposting processes as evaluated by denaturing gradient gel electrophoresisanalysis. J. Appl. Microbiol. 95, 109–119. http://dx.doi.org/10.1046/j.1365-2672.2003.01949.x.

92 I. Villar et al. /Waste Management 54 (2016) 83–92

Jindo, K., Sánchez-Monedero, M.A., Hernández, T., García, C., Furukawa, T.,Matsumoto, K., Sonoki, T., Bastida, F., 2012. Biochar influences the microbialcommunity structure during manure composting with agricultural wastes. Sci.Total Environ. 416, 476–481. http://dx.doi.org/10.1016/j.scitotenv.2011.12.009.

Klamer, M., Bååth, E., 1998. Microbial community dynamics during composting ofstraw material studied using phospholipid fatty acid analysis. FEMS Microbiol.Ecol. 27, 9–20. http://dx.doi.org/10.1016/S0168-6496(98)00051-8.

Klammer, S., Knapp, B., Insam, H., Dell’Abate, M.T., Ros, M., 2008. Bacterialcommunity patterns and thermal analyses of composts of various origins.Waste Manage. Res. 26, 173–187. http://dx.doi.org/10.1177/0734242X07084113.

Ladd, J.N., Butler, J.H.A., 1972. Short-term assays of soil proteolytic enzymeactivities using proteins and dipeptide derivatives as substrates. Soil Biol.Biochem. 4, 19–30. http://dx.doi.org/10.1016/0038-0717(72)90038-7.

Lazcano, C., Gómez-Brandón, M., Domínguez, J., 2008. Comparison of theeffectiveness of composting and vermicomposting for the biologicalstabilization of cattle manure. Chemosphere 72, 1013–1019. http://dx.doi.org/10.1016/j.chemosphere.2008.04.016.

López-González, J.A., Suárez-Estrella, F., Vargas-García, M.C., López, M.J., Jurado, M.M., Moreno, J., 2015. Dynamics of bacterial microbiota during lignocellulosicwaste composting: studies upon its structure, functionality and biodiversity.Bioresour. Technol. 175, 406–416. http://dx.doi.org/10.1016/j.biortech.2014.10.123.

Mondini, C., Fornasier, F., Sinicco, T., 2004. Enzymatic activity as a parameter for thecharacterization of the composting process. Soil Biol. Biochem. 36, 1587–1594.http://dx.doi.org/10.1016/j.soilbio.2004.07.008.

Paradelo, R., Moldes, A.B., Prieto, B., Sandu, R.-G., Barral, M.T., 2010. Can stability andmaturity be evaluated in finished composts from different sources? CompostSci. Util. 18, 22–31. http://dx.doi.org/10.1080/1065657X.2010.10736930.

Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D., R Core Time, 2015. nlme: linear andnonlinear mixed effects models. R package version 3.1-119.

R Development Core Team, 2014. R: a language and environment for statisticalcomputing [WWW Document]. R Found. Stat. Comput. Vienna, Austria. URL<https://www.r-project.org/>.

Ranalli, G., Bottura, G., Taddei, P., Garavani, M., Marchetti, R., Sorlini, C., 2001.Composting of solid and sludge residues from agricultural and food industries.Bioindicators of monitoring and compost maturity. J. Environ. Sci. Heal. 36,415–436. http://dx.doi.org/10.1081/ESE-100103473.

Richardson, M., 2009. The ecology of the zygomycetes and its impact onenvironmental exposure. Clin. Microbiol. Infect. 15, 2–9. http://dx.doi.org/10.1111/j.1469-0691.2009.02972.x.

Riffaldi, R., Levi-Minzi, R., Pera, A., de Bertoldi, M., 1986. Evaluation of compostmaturity by means of chemical and microbial analyses. Waste Manage. Res. 4,387–396. http://dx.doi.org/10.1177/0734242X8600400157.

Ros, M., García, C., Hernández, T., 2006. A full-scale study of treatment of pig slurryby composting: kinetic changes in chemical and microbial properties. WasteManage. 26, 1108–1118. http://dx.doi.org/10.1016/j.wasman.2005.08.008.

Ryckeboer, J., Mergaert, J., Coosemans, J., Deprins, K., Swings, J., 2003a.Microbiological aspects of biowaste during composting in a monitoredcompost bin. J. Appl. Microbiol. 94, 127–137. http://dx.doi.org/10.1046/j.1365-2672.2003.01800.x.

Ryckeboer, J., Mergaert, J., Vaes, K., Klammer, S., De Clercq, D., Coosemans, J., Insam,H., Swings, J., 2003b. A survey of bacteria and fungi occurring duringcomposting and self-heating processes. Ann. Microbiol. 53, 349–410.

Schinner, F., von Mersi, W., 1990. Xylanase-, CM-cellulase- and invertase activity insoil: an improved method. Soil Biol. Biochem. 22, 511–515. http://dx.doi.org/10.1016/0038-0717(90)90187-5.

Shemekite, F., Gómez-Brandón, M., Franke-Whittle, I.H., Praehauser, B., Insam, H.,Assefa, F., 2014. Coffee husk composting: an investigation of the process usingmolecular and non-molecular tools. Waste Manage. 34, 642–652. http://dx.doi.org/10.1016/j.wasman.2013.11.010.

Sims, G.K., Ellsworth, T.R., Mulvaney, R.L., 1995. Microscale determination ofinorganic nitrogen in water and soil extracts. Commun. Soil Sci. Plant Anal. 26,303–316. http://dx.doi.org/10.1080/00103629509369298.

Sims, J.R., Haby, V.A., 1971. Simplified colorimetric determination of soil organicmatter. Soil Sci. 112, 137–141.

Snell-Castro, R., Godon, J.-J., Delgenès, J.-P., Dabert, P., 2005. Characterisation of themicrobial diversity in a pig manure storage pit using small subunit rDNAsequence analysis. FEMS Microbiol. Ecol. 52, 229–242. http://dx.doi.org/10.1016/j.femsec.2004.11.016.

Tiquia, S.M., Tam, N.F.Y., 1998. Composting of spent pig litter in turned and forced-aerated piles. Environ. Pollut. 99, 329–337. http://dx.doi.org/10.1016/S0269-7491(98)00024-4.

Tiquia, S.M., Wan, J.H.C., Tam, N.F.Y., 2002a. Dynamics of yard trimmingscomposting as determined by dehydrogenase activity, ATP content, arginineammonification, and nitrification potential. Process Biochem. 37, 1057–1065.http://dx.doi.org/10.1016/S0032-9592(01)00317-X.

Tiquia, S.M., Wan, J.H.C., Tam, N.F.Y., 2002b. Microbial population dynamics andenzyme activities during composting. Compost Sci. Util. 10, 150–161. http://dx.doi.org/10.1080/1065657X.2002.10702075.

TMECC, 2002. Test Methods for the Examination of Composting and Compost.Composting Council Research and Education Foundation, and US Department ofAgriculture, Bethesda, MD.

Vargas-García, M.C., Suárez-Estrella, F., López, M.J., Moreno, J., 2010. Microbialpopulation dynamics and enzyme activities in composting processes withdifferent starting materials. Waste Manage. 30, 771–778. http://dx.doi.org/10.1016/j.wasman.2009.12.019.

Villar, I., Alves, D., Pérez-Díaz, D., Mato, S., 2016. Changes in microbial dynamicsduring vermicomposting of fresh and composted sewage sludge. WasteManage. 48, 409–417. http://dx.doi.org/10.1016/j.wasman.2015.10.011.

Vu, V.Q., 2011. ggbiplot: A ggplot2 based biplot. R package, version 0.55 [WWWDocument]. URL <https://github.com/vqv/ggbiplot>.

Wagner, M., Loy, A., 2002. Bacterial community composition and function in sewagetreatment systems. Curr. Opin. Biotechnol. 13, 218–227. http://dx.doi.org/10.1016/S0958-1669(02)00315-4.

Zelles, L., 1999. Fatty acid patterns of phospholipids and lipopolysaccharides in thecharacterisation of microbial communities in soil: a review. Biol. Fertil. Soils 29,111–129. http://dx.doi.org/10.1007/s003740050533.

Zucconi, F., Pera, A., Forte, M., de Bertoldi, M., 1981. Evaluating toxicity of immaturecompost. Biocycle 22, 54–57.

CAPÍTULO 2

Seafood-processing sludge composting: changes to

microbial communities and physico-chemical parameters

of static treatment versus for turning during the

maturation stage

Villar I., Alves D., Mato S., 2016. Seafood-Processing Sludge Composting: Changes to

Microbial Communities and Physico-Chemical Parameters of Static Treatment versus

for Turning during the Maturation Stage. PLoS ONE 11(12): e0168590.

https://doi.org/10.1371/journal.pone.0168590

Estado: publicado

Permiso: pdf editor

Capítulo 2

67

RESUMEN

En general, en las plantas de compostaje la fase activa o intensiva se realiza

separadamente de la fase de maduración, usando una tecnología y tiempo específicos. El

material pre-compostado a ser madurado puede presentar suficientes sustratos

biodegradables que provoquen la proliferación microbiana y la consecuente reactivación de

la temperatura. Por lo tanto, la falta de control sobre la estancia en maduración durante la

gestión del residuo a nivel industrial puede resultar en situaciones no deseables. La principal

hipótesis de esta investigación es que el control de la fase de maduración mediante volteos

conlleva una optimización del proceso de compostaje en comparación con una maduración

estática. El residuo empleado fue lodo de depuradora de una industria de procesado de

productos del mar mezclado con madera triturada (1:2, v/v). El sistema de compostaje

consistió en una fase bio-oxidativa en reactor estático de 600L seguido de una fase de

maduración por triplicado en cajas de 200L durante 112 días. Se realizaron dos pruebas con

el mismo proceso en reactor y diferentes tratamientos en cajas: maduración estática y

maduración volteada cuando las temperaturas superaron los 55ºC. Se midieron

periódicamente PLFAs, materia orgánica, pH, conductividad eléctrica, formas de nitrógeno,

formas de carbono, enzimas hidrolíticas y actividad respiratoria. Los volteos aumentaron

significativamente la duración de la fase termófila con el consecuente aumento de la

degradación de la materia orgánica. El PCA diferenció significativamente los dos tratamientos

en función de los parámetros de seguimiento, especialmente, pH, carbono total, formas de

nitrógeno y ratio C/N. En consecuencia, se alcanzaron valores óptimos de madurez y

estabilidad en menor tiempo en el compost con volteos. Mientras que el tratamiento volteado

mostró una estabilización de los grupos microbianos y un bajo ratio mono/sat, el tratamiento

estático presentó mayor variabilidad en los grupos microbianos y un alto ratio mono/sat. La

presencia de sustratos degradables causa cambios en las comunidades microbianas y su

estudio durante la maduración ofrece una aproximación al estado de degradación de la

materia orgánica. La obtención de un compost de calidad y la optimización del proceso

requiere del control mediante volteos durante la etapa de maduración.

RESEARCH ARTICLE

Seafood-Processing Sludge Composting:

Changes to Microbial Communities and

Physico-Chemical Parameters of Static

Treatment versus for Turning during the

Maturation Stage

Iria Villar*, David Alves, Salustiano Mato

Department of Ecology and Animal Biology, University of Vigo, Vigo, Pontevedra, Spain

* [email protected]

Abstract

In general, in composting facilities the active, or intensive, stage of the process is done sep-

arately from the maturation stage, using a specific technology and time. The pre-composted

material to be matured can contain enough biodegradable substrates to cause microbial

proliferation, which in turn can cause temperatures to increase. Therefore, not controlling

the maturation period during waste management at an industrial level can result in unde-

sired outcomes. The main hypothesis of this study is that controlling the maturation stage

through turning provides one with an optimized process when compared to the static

approach. The waste used was sludge from a seafood-processing plant, mixed with shred-

ded wood (1:2, v/v). The composting system consists of an intensive stage in a 600L static

reactor, followed by maturation in triplicate in 200L boxes for 112 days. Two tests were car-

ried out with the same process in reactor and different treatments in boxes: static maturation

and turning during maturation when the temperature went above 55˚C. PLFAs, organic mat-

ter, pH, electrical conductivity, forms of nitrogen and carbon, hydrolytic enzymes and respi-

ratory activity were periodically measured. Turning significantly increased the duration of

the thermophilic phase and consequently increased the organic-matter degradation. PCA

differentiated significantly the two treatments in function of tracking parameters, especially

pH, total carbon, forms of nitrogen and C/N ratio. So, stability and maturity optimum values

for compost were achieved in less time with turnings. Whereas turning resulted in microbial-

group stabilization and a low mono/sat ratio, static treatment produced greater variability in

microbial groups and a high mono/sat ratio, the presence of more degradable substrates

causes changes in microbial communities and their study during maturation gives an

approach of the state of organic-matter degradation. Obtaining quality compost and optimiz-

ing the composting process requires using turning as a control mechanism during

maturation.

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 1 / 15

a1111111111

a1111111111

a1111111111

a1111111111

a1111111111

OPENACCESS

Citation: Villar I, Alves D, Mato S (2016) Seafood-

Processing Sludge Composting: Changes to

Microbial Communities and Physico-Chemical

Parameters of Static Treatment versus for Turning

during the Maturation Stage. PLoS ONE 11(12):

e0168590. doi:10.1371/journal.pone.0168590

Editor: Andrew C Singer, Natural Environment

Research Council, UNITED KINGDOM

Received: June 23, 2016

Accepted: December 2, 2016

Published: December 21, 2016

Copyright: © 2016 Villar et al. This is an open

access article distributed under the terms of the

Creative Commons Attribution License, which

permits unrestricted use, distribution, and

reproduction in any medium, provided the original

author and source are credited.

Data Availability Statement: All relevant data are

within the paper.

Funding: This work was supported by Xunta de

Galicia, 09MDS024310PR. The funders had no role

in study design, data collection and analysis,

decision to publish, or preparation of the

manuscript.

Competing Interests: The authors have declared

that no competing interests exist.

Introduction

Food-processing industries in the European Union generate a large quantity of waste-prod-

ucts, estimated to be 35 megatonnes per year; 60% of this consists of organic material [1]. The

seafood-processing industry generates effluent with a high content of fish remains and oils,

which means wastewater with a high organic load [2]. For the management of organic sludge

from wastewater treatment, one of the most widely-used techniques is composting. Compost-

ing is a process of biologically degrading solid organic substrates under aerobic conditions

through the action of diverse microbial populations; through this process one obtains a stable-

humified-product that is suitable for land application as a soil-improver and source of organic

matter and nutrients [3]. The elevated availability of nourishment in organic waste causes

microbial growth and with it an elevation in temperature and consequently the succession of

different microbial communities which appear at different stages of the composting process.

In the first phase, known as the mesophilic phase, proliferate mesophiles decompose the most

basic organic compounds, causing the temperature to exceed 45˚C. This increase in tempera-

ture results in the growth of thermophilic organisms and the inhibition of thermally-intolerant

ones in what is known as the thermophilic phase. The final phases consist of a cooling period,

which is characterized by the growth of mesophilic organisms, and a maturation period during

which the organic material is stabilized and turned into humus, obtaining a product suitable

for use as a soil amendment [4].

Generally speaking, industrial composting facilities differ in two different areas: the first

one where the material undergoes intensive decomposition (corresponding to the initial meso-

philic and thermophilic phases) and the second one where stabilization and maturation phases

should take place. The intensive stage is characterized by high temperatures, elevated oxygen

consumption, and the production of gaseous and liquid emissions [5] and, therefore, compost-

ing facilities pay special attention to this part of the process, carried out with a specific tech-

nique or technology, with greater control and tracking than the second one. When this stage is

thought to be finished, or the established time of the intensive process ends, the pre-composted

material is deposited in the maturation area where the degradation and polymerization of

organic substances under mesophilic conditions should continue. Nevertheless, the material

obtained after the intensive stage may not be sufficiently decomposed and may undergo reacti-

vation during the maturation stage. For example, in facilities with composting tunnels, the

most active phase is often limited to a specific time (a two-to-three week stay) after which the

content removed from the tunnels is deposited in piles to mature [6]. The limited duration

and the lack of mixture of the waste during the time in the tunnel usually cause the reactivation

of the material in the maturation area and the amount of attention given to this phase depends

on the composting facility, but typically the maturation area is simply considered to be a stor-

age space for compost [7]. The lack of control during this phase can cause environmental

problems such as smells and leachates; it can also lower the quality of the compost. In order to

obtain stable and matured compost, composting-facility operation and design should integrate

both phases [5]. In many cases it is not possible to make changes a posteriori on the technique

used in the intensive stage but it is possible to act on the pre-composted material disposed in

the maturation area. So, it is important to know how the management and control of the mate-

rial deposited in maturation conditions similar to composting facilities affects the physical-

chemical and biological parameters and if this management allows a reduction of time and/or

quality improvement of compost.

Turning is usually employed as a means to aerate, homogenize, and control the temperature

during composting [8]. The effects of turning, along with how often it occurs, have been stud-

ied by a wide range of authors, who have observed changes to physical, chemical, and

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 2 / 15

biological characteristics which affect the rate of decomposition and consequently the time

needed to achieve maturity and high-quality compost [9–12]. Nevertheless, information about

the possible effects of dynamic or static management of the pre-composted waste has not been

found; this is due to the fact that turning is generally carried out during the most active part of

the process, or scheduled throughout the entire process.

Therefore, the intention of this study is to understand, at a pilot scale, the maturation pro-

cess carried out at composting facilities, examining the effects of both static management and

turning on temperature control during this stage of the process. Furthermore, there are few

studies related to seafood-processing sludge composting and even fewer dealing with the mat-

uration phase. In a previous experiment, Villar et al. [13] studied the maturation of this type of

waste product in comparison with pig manure and sludge from an urban wastewater treatment

plant. The research demonstrates that the seafood-processing sludge used contains a high level

of biodegradable substrates, which cause it to reactivate when left to mature after an intensive

period in a static reactor. This made it ideal to be used in this study.

The objectives of this research were, firstly, to study whether static management or turning

for temperature control influence the physico-chemical and biological parameters and the

microbial-community structure throughout the maturation phase and, secondly, to determine

if controlling the maturation phase produced a more stable product in less time. This study

addresses the hypothesis that controlling by turning the pre-composted sludge allows for an

improvement in compost quality, a reduction in the time process to obtain appropriate stabil-

ity and maturity parameters and a microbial-community profile different to the static treat-

ment one.

Materials and Methods

Composting experiment

Composting substrates. For this experiment, the wastewater sludge was from a purifier

from a plant that produces pre-cooked and flash-frozen fish and cephalopods; it was obtained

after fats had been separated from the sludge and the waste product had been treated with

coagulants and flocculants; pinewood shred to smaller than 3cm was used as a bulking agent.

The initial characteristics of these materials can be seen in Table 1. Both materials were mixed

in a volumetric ratio 1:2, respectively, as in the previous study [13].

Composting design. Composting was carried out in two stages, the objective being to

simulate waste-product management as seen in industrial composting facilities. The compost-

ing system has been described in detail in Villar et al. [13]. In brief, the sludge-and-bulking-

Table 1. Physico-chemical composition of the materials used in the composting experiments.

Seafood sludge Bulking agent

Moisture content (%) 61.8 ± 0.4 41.4 ± 0.2

Organic matter (% dw) 87.7 ± 0.3 93.9 ± 0.4

pH 4.90 ± 0.04 6.66 ± 0.01

Electrical conductivity (mS cm-1) 0.55 ± 0.00 0.29 ± 0.01

Total carbon (mg g-1 dw) 513.7 ± 3.2 558.2 ± 2.8

Total nitrogen (mg g-1 dw) 19.24 ± 0.16 12.80 ± 0.29

C/N ratio 26.7 ± 0.2 43.6 ± 0.5

Fat content (% dw) 19.8 ± 0.8 N.D.

dw: dry weight, N.D: not detected

doi:10.1371/journal.pone.0168590.t001

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 3 / 15

agent mixture was composted in a 600-liter-capacity static reactor; aeration was achieved

through programmed forced ventilation of 400 m3/h by a periodic flow rate of 30 seconds

every 90 min and an alarm flow for temperature and oxygen control. The material was kept in

the reactor, where it underwent the intensive stage, until the temperature had dropped to

below 35˚C. The reactor was emptied and the homogenized material placed in triplicate in

200-liter-capacity wooden boxes where the cooling, stabilization, and maturation stage took

place (henceforth called the maturation phase). The top parts of these boxes were open, their

bottom part perforated; inside them, the material rested between two layers of bulking agent to

lend some thermal insulation to the pre-composted material.

Two composting tests were carried out, using the same mixture and handling in the reactor;

each test, however, received a different treatment in boxes, either static treatment or turning

treatment for temperature control. The boxes that received the static treatment were left for

112 days without being homogenized, whereas the ones that received the turning treatment

were mixed when temperatures went above 55˚C to aerate, avoid excess temperature and

homogenize the compost. The turnings were carried out in order to simulate industrial-level

compost turning by emptying completely each box, thoroughly mixing and placing it again

inside the box. The compost was moistened when moisture levels fell below 40% during both

types of treatment. The boxes underwent oxygen and temperature control daily for the first 42

days and three times a week afterwards until the end of the 112-day process. 1500-milliliter of

composite samples were taken from each box on day 0 (when the reactor was emptied) and on

the 14th, 28th, 42nd, 56th, 70th, and 91st days of being in the box; they were screened through

a one-centimeter sieve in order to remove the bulking agent. Once 112 days had passed, the

boxes were emptied and all the compost screened. The samples were divided in three parts.

One part was dried for two days at 40˚C and ground in order to determine the total amount of

carbon and nitrogen. Another part was freeze-dried and ground using mortar-and-pestle in

order to analyze phospholipid fatty acids (PLFAs). The remaining part was kept fresh for the

rest of the analytical determinations.

Composting parameters

Moisture and organic matter contents of samples were calculated gravimetrically after drying

at 105˚C until constant weight and combustion at 550˚C for 4 h, respectively. Total carbon

(TC) and total nitrogen (TN) contents were determined by combustion using a LECO 2000

CN elemental analyzer. Inorganic nitrogen (N-NH4+ and N-NO3

−) was determined in 0.5 M

K2SO4 extracts (1:10, w/v) using the modified indophenol blue colorimetric method [14].

Total extractable nitrogen was determined in the same extracts after oxidation with K2S2O8, as

described by Cabrera and Beare [15], and dissolved organic nitrogen content (DON) was cal-

culated as (total extractable N)–(inorganic N). Water soluble carbon content (WSC) was ana-

lyzed in aqueous extracts (1:5, w/v) by dichromate oxidation in sulfuric acid solution.

Electrical conductivity and pH were determined in aqueous extracts (1:10, w/v) using a pH

meter Crison Basic 20 and a conductivimeter Crison CM 35. Fat content was determined in

samples from day 56 and 112 by Soxhlet method using n-hexane as organic solvent [16].

The microbial community composition and biomass were determined by PLFA analysis

following the method described by Gomez-Brandon et al. [17] for organic samples. The analy-

sis was performed with a CP-Select FAME, 100m x 0.25mm in a gas chromatograph-mass

spectrometer (GC-MS). Identification was done by comparison of retention times and mass

spectra with known external standards (Larodan Fine Chemicals AB, Malmo, Sweden) and

quantification was performed with methyl nonadecanoate fatty acid (C19:0) as internal stan-

dard. PLFAs were used to estimate the biomass of specific microbial groups: Gram-positive

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 4 / 15

bacteria (i14:0, i15:0, a15:0, i16:0, a17:0), Gram-negative bacteria (15:1ω5, 16:1ω7, 17:1ω7,

18:1ω7) and fungi (18:2ω6, 18:1ω9, 20:1ω9). The total amount of PLFAs identified (totPLFAs)

was used as an indicator of the viable microbial biomass [18]. The ratio of monounsaturated

PLFAs to saturated PLFAs (mono/sat) was used as an indicator of physiological or nutritional

stress in microbial communities [19].

β-glucosidase was estimated by the colorimetric measurement of p-nitrophenol released

after incubation 1 g of fresh sample with 1 mL of p-nitrophenyl-β-D-glucopiranoside

(0.025 M) for 1 h at 37˚C [20]. Alkaline phosphatase was measured by incubating 0.5 g of fresh

sample with 1 mL of p-nitrophenylphosphate (0.015 M) for 1 h at 37˚C and subsequent colori-

metric measurement of p-nitrophenol released [21]. Protease activity was measured after the

incubation of 1 g fresh sample with 5 mL of sodium caseinate (2%) for 2 h at 50˚C and subse-

quent determination of the amino acids released using Folin-Ciocalteu reagent [22].

Static respiration rate (SR) was measured using manometric respirometers by OxiTop1

system (WTW GmbH, Weilheim, Germany). Briefly, fresh weight equivalent to 4 g of dry sam-

ple was placed in a hermetic container with a 1 M NaOH trap to capture CO2, the pressure

drop, due to microbial oxygen consumption, was recorded during 24 h at constant tempera-

ture. Germination index (GI) was calculated on day 56 and 112 by determining seed germina-

tion and root length of Lepidium sativum growing in 2 mL of aqueous extracts (1:5, w/v) in

Petri dishes lined with paper filter during 48 h [23]. The self-heating test was carried out in the

final compost using 2 L Dewar flask for 10 days at room temperature.

Statistical analysis

All statistical tests were performed using R software [24]. Principal component analysis (PCA)

was performed on the correlation matrix after normalization to zero mean and variance unit

of the variables. The PCA was conducted using the prcomp function and the package factoex-

tra [25]. Mixed models were fitted with the nlme package [26] to evaluate the differences

between treatments. The waste type and time were fixed factors and the repeated measurement

throughout time in each maturation box was treated as a random effect to address the non-

independence of samples. The best model was selected according to Akaike Information Crite-

rion (AIC). Logarithmic and square root transformations of the data were necessary to ensure

the normality and homogeneity of the variance of residuals of models. Student’s t tests were

performed to determine the difference between compost. PLFAs data of each treatment were

subjected to cluster analysis with the hclust function to determine the differences in the struc-

ture of the microbial community according to time. All statistical tests were evaluated at the

95% confidence level and values are given as the mean ± standard error.

Results

Temperature evolution

The temperature profile during the reactor phase was similar in both tests, as can be seen in

Fig 1; there were no significant differences between them (p> 0.05). The temperature slowly

increased at the start of process, getting above 45˚C starting on the 5th day. Both tests main-

tained thermophilic temperatures for 28 days. After emptying the reactors, reactivation of the

process with thermophilic temperatures was quickly reached in the box tests.

The temperature evolutions of the two treatments during maturation were significantly dif-

ferent (p< 0.05). Static treatment saw temperatures maintained constantly in excess of 45˚C

for ten days, with a second peak temperature being registered on the 23rd day at no higher

than 45˚C; on the 14th day it was necessary to remoisten the material in order to keep moisture

levels above 40%. Turning treatment saw the material in the boxes being turned seven times

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 5 / 15

over a period of 20 days; the material was remoistened during three of these turns and thermo-

philic temperatures were maintained for 26 days. Both treatments saw oxygen levels superior

to 14% during the material’s time in boxes.

Composting parameters

The PCA was undertaken with the physico-chemical and biological variables shown in Fig 2.

Three separate groups were yielded; they were the one formed at the start of the maturation

phase of both tests, in other words when the materials left the reactor (time 0), the turning-

treatment sample group, and the static-treatment sample group. The principal component

1 (PC1) accounted for 51.2% of the variance and significantly separated the three groups

(F2, 45 = 337.95, p< 0.0001). The parameters responsible for the differences between the

groups through PC1 were, fundamentally, pH (r = 0.96, p<0.0001), TC (r = -0.94, p<0.0001),

TN (r = 0.92, p<0.0001), and the C/N ratio (r = -0.97, p<0.0001).

The parameters that contributed the most to separating the groups during the principal

component 2 (PC2), which accounted for 29% of the variance, were DON (r = 0.88, p

<0.0001), and ammonium (r = 0.90, p<0.0001); the three groups were significantly different

from each other (F2, 45 = 29.09, p<0.0001). The initial material of the maturation stage con-

tained high amounts of TC, OM, SR, and totPLFAs and low amounts of pH, TN, β-glucosidase

and protease activity. Static treatment was characterized by higher levels of ammonium

(p<0.0001), EC (p<0.0001), DON (p<0.0001), OM (p<0.0001), pH (p<0.0001), alkaline

phosphatase (p<0.0001), protease (p<0.0001), and TN (p<0.0001) during the maturation

process than those obtained via the turning treatment; static treatment produced lower levels

Fig 1. Temperature evolution during the reactor stage and the maturation stage for the turning treatment and static treatment.

doi:10.1371/journal.pone.0168590.g001

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 6 / 15

of β-glucosidase (p<0.0001) than that achieved via turning. Organic-matter-loss at the end of

maturation during the box phase was greater in the case of the turning treatment (42.5%) than

the static one (35.6%); likewise, the percentage of carbon reduction was greater in the turned

boxes (17.4%) than in the static ones (10.6%). Both treatments presented significantly different

parameters in the middle of the process (56th day); these differences, with the exception of the

C/N ratio, which was present at similar levels in both types of compost (Table 2), were main-

tained up to, and including, the final sampling.

Fig 2. A) Principal component analysis (PCA) of the physico-chemical and biological variables of the turning treatment and static treatment B)

Correlation circle showing the variables that define the components. CN: carbon-to-nitrogen ratio, DON: dissolved organic nitrogen, EC: electrical

conductivity, OM: organic matter, SR: static respiration, TC: total carbon, TN: total nitrogen, totPLFAs: total amount of PLFAs, WSC: water soluble carbon.

doi:10.1371/journal.pone.0168590.g002

Table 2. Compost characteristics after 56 and 112 days of maturation in the turning treatment and static treatment.

56 days 112 days

Turning Static Turning Static

Organic Matter (%) 74.3 ± 0.3 76.8± 0.8* 70.1 ± 0.5 75.7 ± 0.5*

pH (mS cm-1) 5.99 ± 0.03 6.61± 0.03* 6.01 ± 0.04 5.98 ± 0.05

ratio C/N 15.1 ± 0.5 12.7± 0.2* 13.8 ± 0.2 13.1 ± 0.2

NH4+/NO3

- 0.17 ± 0.02 0.23 ± 0.01* 0.11 ± 0.01 0.21 ± 0.01*

SR (mg O2 g-1OM h-1) 0.48 ± 0.03 0.64 ± 0.02* 0.36 ± 0.02 0.47± 0.02*

Sel-heating test - - clase V clase V

GI(%) 82.7 ± 1.5 58.3 ± 2.1* 122.9± 6.9 77.3 ± 5.7*

Fat (%) 6.03 ± 0.03 9.14 ± 0.03* 4.21 ± 0.02 8.00 ± 0.01*

* indicates that samples for the same time between treatments are significantly different (Student t-test, p < 0.05) (SR: static respiration, GI: germination

index).

doi:10.1371/journal.pone.0168590.t002

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 7 / 15

Microbial community

Cluster analysis based on PLFA profiles for turned boxes and static boxes show two clusters

for each of the treatments. In the case of the turned boxes (Fig 3A), the samples on day 0, the

14th day, and the 28th day formed a cluster that was separate from the rest of the samplings,

although the distance between day 0, the 14th day, and the 28th day was high; this suggests

that the PLFA profiles were different. In the case of the second group (the 42nd day, 56th day,

72nd day, 91st day, and the 112th day), the similarities between the samples within the cluster

were greater, with little distance between the samplings. With regards to static boxes (Fig 3B),

the samples on day 0 and the 14th day formed a cluster, although the distance between one

another was high; this suggests important differences between both samplings. Significant

increases in bond distance were observed in the second cluster (the 28th day, 42nd day, 56th

day, 70th day, 91st day, and the 112th day), indicating a low level of similarity between the

samplings, especially between the 28th-day sampling and the rest, and between 70th-day-to-

112th-day subgroup and the 56th-day-and-42nd-day-and-91st-day subgroup.

Table 3 shows the evolution of microbial groups during their time in boxes. The

treatment performed on the material of the boxes significantly affected microbial-biomass evo-

lution (p<0.0001), Gram + bacteria (p<0.0001), Gram—bacteria (p<0.0001), and fungi

(p<0.0001). Likewise, significant differences caused by time (p<0.0001) and significant inter-

action between time and treatment (p<0.0001) were observed in the microbial biomass and

all microbial groups. Both maturation treatments showed a high concentration of PLFA fungal

biomarkers for the first samplings, with a decrease over time; this was especially true for the

turning treatment (88.9% reduction in the case of turning versus 60.7% in the case of static

treatment). Fungi were the main microbial group (averaging 48%) for practically the entire

process, with the exception of the 28th-day-and-42nd-day-turning-treatment samplings in

which Gram + bacteria were dominant. Gram + bacteria were reduced by 81% and 61% during

the turning treatment and static treatment, respectively. The PLFA-Gram-negative-bacteria

Fig 3. Dendrograms of cluster analysis based on PLFA profiles during the maturation stage in boxes of A) turning treatment and B) static

treatment. Dendrograms were done with the data of 8 samplings taken during the maturation stage (day 0 to the 112th day) and were drawn based on

Ward’s method on the Euclidean distance.

doi:10.1371/journal.pone.0168590.g003

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 8 / 15

indicators showed a high increase on the 28th day of the static-treatment process; much higher

levels than the turning treatment were maintained until the end of the process; static treatment

showed an increase of 33%, whereas the turning treatment showed a 74% reduction. The com-

post produced by static treatment had a higher-level concentration of all the PLFA-group indi-

cators than the turned compost.

The mono/sat ratio achieved by static treatment was greater than that achieved by turning

starting from the 28th day, the latter maintaining stable levels starting from the 70th day.

Discussion

The temperature profiles of both tests while in the reactor, corresponding to the intensive

stage, showed patterns typical of the composting process; in other words, temperatures

increasing to thermophilic levels followed by maintaining said temperature and a subsequent

decline in temperature until reaching mesophilic levels. Despite thermophilic conditions being

reached slowly, compost hygienization was ensured by continuously maintaining tempera-

tures above 55˚C for more than 15 days [27]. Similar temperature profiles have been observed

by other authors during the composting of fatty wastes [8,28,29], therefore, a high level of fat

causes the compost to reach thermophilic temperatures more slowly but once reached they

cause the compost to maintain thermophilic temperatures for longer due to the fact that the

lipids present provide a greater amount of energy than other organic compounds. The sludge

used in the research came from the purification of wastewater generated by the production of

precooked and frozen foods, such as breaded and battered fish and cephalopods, so the lipid

content of the waste was high (Table 2). Thus, the sludge used had an optimum amount of fats

for microbial degradation through composting since the appropriate values for this process

should not exceed 20–25% according to studies carried out by Fernandes et al. [28] for urban,

agricultural and industrial waste. Likewise, the low pH of the sludge, with values less than 5,

was able to inhibit thermophilic microorganisms and slow the transition from mesophilic to

thermophilic temperatures during the initial composting reactor stage, as observed by Sund-

berg et al. [30]. The lack of difference between the temperature profiles of both tests in the

reactor makes it possible to carry out statistical analysis conducive to checking the current

study’s objectives by submitting the material to the same processing conditions during the

intensive phase.

After the static-reactor stage, the pre-composted material was homogenized and placed in

boxes, resulting in a quick reactivation for both treatments, indicating that the material had

Table 3. Changes in microbial groups: bacteria Gram +, bacteria Gram—and fungi and ratio of monounsaturated to saturated (mono/sat) PLFAs

during the maturation stage of the turning treatment and static treatment.

Time (days) Gram + (μg g-1dw) Gram–(μg g-1dw) Fungi (μg g-1dw) mono/sat

Turning Static Turning Static Turning Static Turning Static

0 139.9 ± 3.9 176.1 ± 8.9* 44.5 ± 4.1 39.0 ± 1.3 265.5 ± 2.1 276.4 ± 14.9 1.61 ± 0.10 1.25 ± 0.01*

14 119.8 ± 4.3 117.8 ± 5.7 20.0 ± 1.2 22.5 ± 1.4 244.9 ± 12.0 147.6 ± 10.3* 1.32 ± 0.07 0.99 ± 0.02*

28 122.4 ± 7.7 91.3 ± 1.9* 23.4 ± 1.6 90.6 ± 5.5* 109.5 ± 6.6 106.8 ± 8.5 1.04 ± 0.06 1.50 ± 0.09*

42 41.0 ± 5.9 57.5 ± 3.1 13.3 ± 0.8 52.8 ± 6.6* 29.8 ± 3.1 76.6 ± 8.3* 0.78 ± 0.06 1.22 ± 0.06*

56 26.4 ± 2.1 60.2 ± 3.9* 8.1 ± 1.0 52.9 ± 5.4* 27.4 ± 3.5 79.2 ± 3.8* 0.59 ± 0.09 1.27 ± 0.10*

70 39.4 ± 3.1 37.1 ± 1.9 10.4 ± 2.3 26.5 ± 3.8* 45.6 ± 4.9 76.3 ± 7.0* 0.78 ± 0.06 1.11 ± 0.06*

91 41.2 ± 4.2 61.4 ± 2.0* 14.5 ± 1.0 46.1 ± 4.3* 45.6 ± 2.3 94.7 ± 8.7* 0.83 ± 0.05 1.18 ± 0.05*

112 26.6 ± 2.0 68.5 ± 4.5* 11.7 ± 0.7 25.8 ± 0.4* 29.4 ± 0.6 108.7 ± 6.9* 0.79 ± 0.03 1.38 ± 0.06*

* indicates that samples for the same time between treatments are significantly different (Student t-test, p < 0.05) dw: dry weight

doi:10.1371/journal.pone.0168590.t003

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 9 / 15

enough biodegradable substrates to allow it to self-heat until it reached thermophilic tempera-

tures. Likewise, turnings carried out during the first several days of the material being in the

boxes facilitated the extension of thermophilic conditions more so than the static approach.

Therefore, turning reactivated the process of incorporating the least-degraded material and, as

a result, made it possible to supply easily assimilable substrates to the microbial biomass, start-

ing from the exterior of the boxes and moving inwards. Ruggieri et al. [8] have suggested that

turning during the composting of fatty wastes is preferable to the static system because it pre-

vents the mixture from forming clumps and compacting. Likewise, Albuquerque et al. [31]

have observed that, during alperujo composting, forced-air ventilation is only effective when

done along with turning because it improves the substrate’s porosity. It has also showed that

turning, after static-reactor composting seafood sludge with forced ventilation, makes length-

ening the thermophilic phase possible. At most composting facilities, turnings of the pre-

composted waste in the maturation stage are not based on the biological process, but on indus-

trial criteria (increase space, move the material to another location, dry the material. . .) so, the

implementation of protocols for the maturation stage is considered important to improve the

composting process as has been observed.

According to multivariate analysis, the treatment applied while the material is in the boxes

determines its physico-chemical, and biological, development vis-à-vis compost stabilization

and maturation. Turning the material during the maturation stage allows for compost to

degrade to a greater degree, suggesting that the time needed to achieve stability is reduced

when controlling the process is done under dynamic conditions. Other authors have found

similar results with regards to all stages of the composting process and not only the maturation

stage in particular. For example, Brito et al. [32] have found that turning increases the degrada-

tion rate during the composting of cattle slurry; statically-treated piles require more processing

time to reach organic-material values similar to that obtained in compost that had been

turned. Nikaeen et al. [33] have also pointed out that the time needed to stabilize organic com-

pounds in static piles is greater due to the lack of heat exchange. Furthermore, it has been

stated that turning and how often it occurs affects the EC, pH, TC, TN, C/N ratio, GI, and tem-

perature during the composting of different types of waste products [9–12,34]. Similar results

have been obtained from this study, showing that turning pre-composted waste in line with

temperature criteria following static-reactor composting affected the pH, electrical conductiv-

ity, ammoniacal nitrogen, dissolved organic nitrogen, organic matter, enzymatic activities,

total nitrogen and total carbon; this led to the compost becoming stable in less time than

through static conditions for the maturation stage. Turning-treatment samples starting from

the 56th day showed stable conditions; some parameters, like the respiration rate and the ger-

mination index for example, were improved by prolonging the process. Nonetheless, the stati-

cally-treated compost had insufficient quality parameters on the 56th day; there were also

significant variations throughout the entire process. The analysis of the compost after 112 days

in the maturation boxes showed parameters that indicated stability in both cases; these param-

eters included the maximum rating for the Dewar self-heating test (class V, mature compost)

[35], and a respiration rate that was less than 0.5 mgO2 g-1SV h-1 [36]. Also, both types of treat-

ment produced compost that reached optimum C/N-ratio levels of less than 20 [35]. However,

the turned compost had an ammonium/nitrate ratio of less than 0.16 [37] and a germination

index superior to 80% [38], which showed a greater degree of maturity than the compost that

received the static treatment. Controlling through turning significantly affected physico-

chemical, and biological, characteristics as well as microbial activity; making it possible to

achieve a more stable, and more mature, product in less time than the unhandled material. So,

when a composting facility is decided to work with a static technique for the intensive stage,

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 10 / 15

the control of the maturation by turnings could optimize the process to obtain a compost of

better quality in less time.

It is known that the nature of organic substrates and temperature are the principal factors

that determine the composting process with respect to microbial dynamics [4,39,40]. Due to

the fact that the initial material placed in boxes after having been composted in the reactor

were very similar on a physico-chemical, and biological, level -as observed using multivariate

analysis- the changes in the microbial community’s structure were mainly produced as a con-

sequence of the temperature reached in the boxes during the first weeks of maturation stage.

Bacterial and fungal populations decreased with both treatments, which is an expected decline

as the most easily assimilable substrates are consumed, reducing the available food for micro-

bial growth. Nevertheless, turning produced a less noticeable decrease of Gram + bacteria dur-

ing the first few weeks, which was due to Gram + bacteria, in particular those that belong to

the genus Bacillus, dominating during the thermophilic phase of composting [4] and turning

boxes maintaining thermophilic temperatures for more days. In the same way, it is known that

increasing temperatures caused fungi to decrease [4,41], although in the case of turning the

population stayed the same for the first month when the temperatures were high and this

could mean that were thermophilic or thermo-tolerant fungi present [13,42]. Choi and Park

[43] have also observed high activity of yeast in thermophilic composting of food waste, indi-

cating that the ability of yeast to grow at a lower pH than bacteria could explain their presence.

The pH and the lipid content of the pre-composted sludge allowed for the development of

thermophilic fungi and turnings allowed for the input of food to initially maintain their popu-

lation. So, the cluster obtained by analyzing the PLFAs found in the first two static-treatment

samplings, and those obtained from the first three turning-treatment samplings, are mainly

characterized by the temperature reached during the first few weeks which in turn depended

on the treatment carried out.

It is normal for the number of bacteria to lower in population, but increase in terms of

diversity, and for fungi to increase both in terms of quantity and diversity during the matura-

tion phase itself [4]. Nonetheless, a decrease in temperature did not produce an observable

increase in how abundant fungi were; both treatments, however, were characterized by the

fungal biomass being predominant, particularly in the case of static treatment, throughout the

entire process. Amir et al. [44] found that fungi are more present in wastes with a high level of

fatty acids. Likewise, Villar et al. [13] suggested that the influence the initial material, in this

case lipid in nature and pH, has on the maturation phase is reflected by microbial diversity;

this means that although the waste undergoes significant degradation during the composting

process, the properties of the starting material determine the dominant microbial groups

throughout the maturation stage. Following the high-temperature phase, there was a large

increase in Gram-negative-bacteria PLFA biomarkers with respect to static treatment. Due to

Gram—bacteria’s limited resistance to temperature the normal course of action is for thermo-

philic Gram+ bacteria to give way to mesophilic Gram—bacteria during composting or matu-

ration [45]. It has been confirmed that bacteria belonging to the phylum Bacteroidetes and

class Alphaproteobacteria frequently dominate both in compost after it has reached high tem-

peratures and in compost that has not matured [46–48]. Likewise, other Gram—bacteria like

Gammaproteobacteria are dominant in cured compost [48]. However, the fluctuation of

Gram—bacteria together with the physico-chemical data could indicate a lack of stability in

the static treatment in the final stages of maturation. On the contrary, all the microbial groups

in the turned compost stayed at similar levels following the high-temperature phase; this pro-

duced microbial-community homogeneity as observed in group II of the multivariate analysis.

Microbial-population values in turned compost stayed low once it had reached stability and

maturation parameters, whereas the increase in Gram + bacteria and fungi obtained from

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 11 / 15

static treatment could indicate the availability of biodegradable substrates [49]. Turning the

material made it possible to maintain more stable and more similar microbial diversity after

reaching mesophilic temperatures whereas not homogenizing the boxes maintained a greater

degree of difference between samplings. Cahyani et al. [46] have pointed out that the microbial

community remained stable during the maturation phase of composting rice straw. On the

other hand, Danon et al. [48] observed changes to the microbial communities during the year-

long composting of biosolids. These authors observed that in the case of bacterial populations

specialized in breaking down macromolecules such as lignocellulose, stability was maintained

during the intermediate part of the maturation phase.

Likewise, the mono/sat ratio values after high temperatures in the static treatment indicated

that microbial communities do not have limited resources and there are enough nutrients for

their growth. However, in the compost that received the turning treatment the mono/sat ratio

showed that the microbial communities were subjected to nutritional and physiological stress.

Jindo et al. [50] have suggested that low mono/sat ratio numbers demonstrates the final part of

organic material’s transformation into compost and Bastida et al. [51] have pointed to a greater

degree of carbohydrate-substrate biodegradability when the numbers of the aforementioned

ratio are high. Therefore, mono/sat-ratio stabilization can be indicative of mature compost.

Bacterial-population stabilization, as a result of running out of easily degradable substrates,

can indicate the degradation of more calcitrant compounds, ergo turning the material while it

is in maturation stage allows for a greater degree of organic-material degradation and micro-

bial succession towards a more stable mesophilic community, reflecting compost maturation.

Conclusions

The fresh-compost-turning system made it possible to maintain temperatures and, as a result,

prolonged the thermophilic phase; this allowed for a high level of organic-matter degradation

over a longer period. Controlling the composting of highly-energy- waste through turning

after the most intensive stage in the reactor made it possible to achieve stability and maturity

in a shorter time frame than statically treating pre-composted waste. Studying microbial dyna-

mism has helped to characterize the state of degradation of organic matter as stable compost

showing a stable microbial structure whereas the presence of biodegradable substrates causes

significant changes to microbial populations. Looking at the results, composting plants that

utilize turning during the maturation period optimize the process by cutting the time needed

to reach sufficiently high levels of quality.

Acknowledgments

The authors thank the research support services of the University of Vigo (CACTI) for the car-

bon and nitrogen analysis. The authors also thank Emilio Rodrıguez Cochon, Domingo Perez

Dıaz and Josefina Garrido Gonzalez for their help and technical support.

Author Contributions

Conceptualization: SM.

Data curation: IV.

Formal analysis: IV DA.

Funding acquisition: SM.

Investigation: IV DA SM.

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 12 / 15

Methodology: IV SM.

Project administration: SM.

Resources: SM.

Supervision: SM.

Writing – original draft: IV DA SM.

Writing – review & editing: IV DA.

References1. Lin CSK, Pfaltzgraff LA, Herrero-Davila L, Mubofu EB, Abderrahim S, Clark JH, et al. Food waste as a

valuable resource for the production of chemicals, materials and fuels. Current situation and global per-

spective. Energy Environ Sci. 2013; 6: 426–464.

2. Garcıa-Morales JL, Alvarez CJ, Paredes C, Lopez E, Fernandez FJ, Bustamante MA, et al. Residuos

agroalimentarios I.3. Moreno J, Moral R, Garcıa-Morales JL, Pascual JA, Bernal MP, editors. Madrid:

Mundi-Prensa; 2015.

3. Insam H, de Bertoldi M. Microbiology of the composting process. In: Diaz LF, de Bertoldi M, Bidlingma-

ier W, Stentinford E, editors. Compost Science and Technology Waste Management Series. Elsevier

Ltd.; 2007. pp. 25–48.

4. Ryckeboer J, Mergaert J, Vaes K, Klammer S, De Clercq D, Coosemans J, et al. A survey of bacteria

and fungi occurring during composting and self-heating processes. Ann Microbiol. 2003; 53: 349–410.

5. Haug RT. The Practical Handbook of Compost Engineering. Boca Raton: Lewis Publishers; 1993.

6. Diaz LF, Savage GM, Golueke CG. Composting of municipal solid wastes. In: Tchobanoglous G, Kreith

F, editors. Handbook of solid waste management. New York: McGraw-Hill Inc.; 2002. pp. 12.1–12.70.

7. Rynk R. Fires At Composting Facilities: Causes And Conditions. Biocycle. JG Press, Inc.; 2000; 41: 54.

8. Ruggieri L, Artola A, Gea T, Sanchez A. Biodegradation of animal fats in a co-composting process with

wastewater sludge. Int Biodeterior Biodegrad. 2008; 62: 297–303.

9. Tiquia SM, Tam NFY, Hodgkiss IJ. Effects of turning frequency on composting of spent pig-manure

sawdust litter. Environ Pollut. 1997; 62: 37–42.

10. Wong JWC, Mak KF, Chan NW, Lam A, Fang M, Zhou LX, et al. Co-composting of soybean residues

and leaves in Hong Kong. Bioresour Technol. 2001; 76: 99–106. PMID: 11131806

11. Ogunwande GA, Osunade JA, Adekalu KO, Ogunjimi LAO. Nitrogen loss in chicken litter compost as

affected by carbon to nitrogen ratio and turning frequency. Bioresour Technol. 2008; 99: 7495–503. doi:

10.1016/j.biortech.2008.02.020 PMID: 18367393

12. Getahun T, Nigusie A, Entele T, Van Gerven T, Van der Bruggen B. Effect of turning frequencies on

composting biodegradable municipal solid waste quality. Resour Conserv Recycl. 2012; 65: 79–84.

13. Villar I, Alves D, Garrido J, Mato S. Evolution of microbial dynamics during the maturation phase of the

composting of different types of waste. Waste Manag. 2016; 54: 83–92. doi: 10.1016/j.wasman.2016.

05.011 PMID: 27236404

14. Sims GK, Ellsworth TR, Mulvaney RL. Microscale determination of inorganic nitrogen in water and soil

extracts. Commun Soil Sci Plant Anal. 1995; 26: 303–316.

15. Cabrera ML, Beare MH. Alkaline persulfate oxidation for determining total nitrogen in microbial biomass

extracts. Soil Sci Soc Am J. Soil Science Society of America; 1993; 57: 1007–1012.

16. APHA, AWWA, WEF. Standard Methods for the Examination of Water and Wastewater. Stand Meth-

ods. 2012;

17. Gomez-Brandon M, Lores M, Domınguez J. A new combination of extraction and derivatization meth-

ods that reduces the complexity and preparation time in determining phospholipid fatty acids in solid

environmental samples. Bioresour Technol. 2010; 101: 1348–1354. doi: 10.1016/j.biortech.2009.09.

047 PMID: 19800785

18. Zelles L. Fatty acid patterns of phospholipids and lipopolysaccharides in the characterisation of micro-

bial communities in soil: a review. Biol Fertil Soils. 1999; 29: 111–129.

19. Bossio DA, Scow KM. Impacts of carbon and flooding on soil microbial communities: phospholipid fatty

acid profiles and substrate utilization patterns. Microb Ecol. 1998; 35: 265–278. PMID: 9569284

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 13 / 15

20. Eivazi F, Tabatabai MA. Glucosidases and galactosidases in soils. Soil Biol Biochem. 1988; 20:

601–606.

21. Eivazi F, Tabatabai MA. Phosphatases in soils. Soil Biol Biochem. 1977; 9: 167–172.

22. Ladd JN, Butler JHA. Short-term assays of soil proteolytic enzyme activities using proteins and dipep-

tide derivatives as substrates. Soil Biol Biochem. 1972; 4: 19–30.

23. Zucconi F, Pera A, Forte M, de Bertoldi M. Evaluating toxicity of immature compost. Biocycle. 1981; 22:

54–57.

24. R Development Core Team. R: a language and environment for statistical computing [Internet]. R

Foundation for Statistical Computing, Vienna, Austria. 2014. https://www.r-project.org/

25. Kassambara A. factoextra: Visualization of the outputs of a multivariate analysis [Internet]. R package

version 1.0.1. 2015. https://cran.r-project.org/web/packages/factoextra/index.html

26. Pinheiro J, Bates D, DebRoy S, Sarkar D, R Core Time. nlme: linear and nonlinear mixed effects mod-

els. R package version 3.1–119 [Internet]. 2015. https://cran.r-project.org/web/packages/nlme/nlme.pdf

27. European Commission. Working document: biological treatment of biowaste, 2nd draft. Directorate-

General Environment. Brussels; 2001.

28. Fernandes F, Viel M, Sayag D, Andre L. Microbial breakdown of fats through in-vessel co-composting

of agricultural and urban wastes. Biol Wastes. 1988; 26: 33–48. http://dx.doi.org/10.1016/0269-7483

(88)90147-4

29. Gea T, Ferrer P, Alvaro G, Valero F, Artola A, Sanchez A. Co-composting of sewage sludge: fats mix-

tures and characteristics of the lipases involved. Biochem Eng J. 2007; 33: 275–283.

30. Sundberg C, Smårs S, Jonsson H. Low pH as an inhibiting factor in the transition from mesophilic to

thermophilic phase in composting. Bioresour Technol. 2004; 95: 145–150. doi: 10.1016/j.biortech.2004.

01.016 PMID: 15246438

31. Alburquerque JA, Gonzalvez J, Tortosa G, Baddi GA, Cegarra J. Evaluation of “alperujo” composting

based on organic matter degradation, humification and compost quality. Biodegradation. 2009; 20:

257–70. doi: 10.1007/s10532-008-9218-y PMID: 18814039

32. Brito LM, Coutinho J, Smith SR. Methods to improve the composting process of the solid fraction of

dairy cattle slurry. Bioresour Technol. 2008; 99: 8955–60. doi: 10.1016/j.biortech.2008.05.005 PMID:

18556195

33. Nikaeen M, Nafez AH, Bina B, Nabavi BF, Hassanzadeh A. Respiration and enzymatic activities as indi-

cators of stabilization of sewage sludge composting. Waste Manag. 2015; 39: 104–110. doi: 10.1016/j.

wasman.2015.01.028 PMID: 25728091

34. Cook KL, Ritchey EL, Loughrin JH, Haley M, Sistani KR, Bolster CH. Effect of turning frequency and

season on composting materials from swine high-rise facilities. Waste Manag. 2015;

35. TMECC. Test Methods for the Examination of Composting and Compost. Thompson WH, Leege PB,

Millner PD, Watson ME, editors. Bethesda, MD: Composting Council Research and Education Foun-

dation, and US Department of Agriculture; 2002.

36. Iannotti DA, Pang T, Toth BL, Elwell DL, Keener HM, Hoitink HAJ. A quantitative respirometric method

for monitoring compost stability. Compost Sci Util. 1993; 1: 52–65.

37. Bernal MP, Paredes C, Sanchez-Monedero MA, Cegarra J. Maturity and stability parameters of com-

posts prepared with a wide range of organic wastes. Bioresour Technol. 1998; 63: 91–99.

38. Zucconi F, Monaco A, Forte M, Bertoldi M. Phytotoxins during the stabilization of organic matter. In:

Gasser JK., editor. Composting of agricultural and other wastes. London: Elsevier Applied Science

Publisher; 1985. pp. 73–85.

39. McKinley VL, Vestal JR. Physical and chemical correlates of microbial activity and biomass in compost-

ing municipal sewage sludge. Appl Environ Microbiol. 1985; 50: 1395–403. PMID: 16346940

40. Vargas-Garcıa MC, Suarez-Estrella F, Lopez MJ, Moreno J. Microbial population dynamics and

enzyme activities in composting processes with different starting materials. Waste Manag. 2010; 30:

771–778. doi: 10.1016/j.wasman.2009.12.019 PMID: 20096556

41. Klamer M, Bååth E. Microbial community dynamics during composting of straw material studied using

phospholipid fatty acid analysis. FEMS Microbiol Ecol. 1998; 27: 9–20.

42. Tuomela M, Vikman M, Hatakka A, Itavaara M. Biodegradation of lignin in a compost environment: A

review. Bioresour Technol. 2000; 72: 169–183.

43. Choi MH, Park YH. The influence of yeast on thermophilic composting of food waste. Lett Appl Micro-

biol. 1998; 26: 175–178. PMID: 9569704

44. Amir S, Abouelwafa R, Meddich A, Souabi S, Winterton P, Merlina G, et al. PLFAs of the microbial com-

munities in composting mixtures of agro-industry sludge with different proportions of household waste.

Int Biodeterior Biodegradation. 2010; 64: 614–621.

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 14 / 15

45. Boulter-Bitzer JI, Trevors JT, Boland GJ. A polyphasic approach for assessing maturity and stability in

compost intended for suppression of plant pathogens. Appl Soil Ecol. 2006; 34: 65–81.

46. Cahyani VR, Matsuya K, Asakawa S, Kimura M. Succession and phylogenetic composition of bacterial

communities responsible for the composting process of rice straw estimated by PCR-DGGE analysis.

Soil Sci Plant Nutr. 2003; 49: 619–630.

47. Green SJ, Michel FC, Hadar Y, Minz D. Similarity of bacterial communities in sawdust- and straw-

amended cow manure composts. FEMS Microbiol Lett. 2004; 233: 115–123. doi: 10.1016/j.femsle.

2004.01.049 PMID: 15043877

48. Danon M, Franke-Whittle IH, Insam H, Chen Y, Hadar Y. Molecular analysis of bacterial community suc-

cession during prolonged compost curing. FEMS Microbiol Ecol. 2008; 65: 133–144. doi: 10.1111/j.

1574-6941.2008.00506.x PMID: 18537836

49. Elouaqoudi FZ, El Fels L, Amir S, Merlina G, Meddich A, Lemee L, et al. Lipid signature of the microbial

community structure during composting of date palm waste alone or mixed with couch grass clippings.

Int Biodeterior Biodegradation. 2015; 97: 75–84.

50. Jindo K, Sanchez-Monedero MA, Hernandez T, Garcıa C, Furukawa T, Matsumoto K, et al. Biochar

influences the microbial community structure during manure composting with agricultural wastes. Sci

Total Environ. 2012; 416: 476–481. doi: 10.1016/j.scitotenv.2011.12.009 PMID: 22226394

51. Bastida F, Kandeler E, Moreno JL, Ros M, Garcıa C, Hernandez T. Application of fresh and composted

organic wastes modifies structure, size and activity of soil microbial community under semiarid climate.

Appl Soil Ecol. 2008; 40: 318–329.

Seafood-Processing Sludge Composting: Static Treatment versus Turning during the Maturation Stage

PLOS ONE | DOI:10.1371/journal.pone.0168590 December 21, 2016 15 / 15

CAPÍTULO 3

Changes in microbial dynamics during vermicomposting of

fresh and composted sewage sludge

Villar, I., Alves, D., Pérez-Díaz, D., Mato, S., 2016. Changes in microbial dynamics

during vermicomposting of fresh and composted sewage sludge. Waste Manag. 48,

409–417. doi:10.1016/j.wasman.2015.10.011

Estado: publicado

Permiso: post print

Capítulo 3

87

RESUMEN

El lodo de depuradora municipal es un residuo con alta carga orgánica que se genera en

grandes cantidades y el cual puede ser tratado mediante técnicas de biodegradación para

reducir su riesgo para el medio ambiente. Esta investigación estudia el vermicompostaje y el

vermicompostaje después del compostaje de lodo de depuradora de aguas residuales con la

especie Eisenia andrei. Para determinar el efecto que las lombrices ejercen sobre la dinámica

microbiana en función del tratamiento, se evaluó la estructura y actividad de la comunidad

microbiana, mediante el análisis de ácidos grasos fosfolípidos y las actividades enzimáticas,

durante 112 días de vermicompostaje de lodo de depuradora fresco y compostado con y sin

lombrices de tierra. La presencia de las lombrices de tierra redujo significativamente la

biomas microbiana y todos los grupos microbianos (bacterias Gram +, bacterias Gram – y

hongos), así como las actividades celulasa y fosfatasa ácida. El tratamiento combinado

compostaje-vermicompostaje mostró un menor desarrollo de lombrices, una mayor biomasa

bacteriana y fúngica que el tratamiento de vermicompostaje, y mayores diferencias

comparado con el control sin lombrices en celulasa, β-glucosidasa, fosfatasa alcalina y ácida.

Ambos tratamientos son adecuados para la estabilización de lodo de depuradora y el

tratamiento combinado compostaje-vermicompostaje puede ser un proceso viable para la

maduración del compost fresco.

Palabras claves: actividades enzimáticas, PLFAs, comunidad microbiana, lombrices de

tierra, residuo orgánico, estabilidad

Capítulo 3

89

Capítulo 3

90

ABSTRACT

Municipal sewage sludge is a waste with high organic load generated in large quantities

that can be treated by biodegradation techniques to reduce its risk to the environment. This

research studies vermicomposting and vermicomposting after composting of sewage sludge

with the earthworm specie Eisenia andrei. In order to determine the effect that earthworms

cause on the microbial dynamics depending on the treatment, the structure and activity of the

microbial community was assessed using phospholipid fatty acid analysis and enzyme

activities, during 112 days of vermicomposting of fresh and composted sewage sludge, with

and without earthworms. The presence of earthworms significantly reduced microbial

biomass and all microbial groups (Gram + bacteria, Gram – bacteria and fungi), as well as

cellulase and alkaline phosphatase activities. Combined composting-vermicomposting

treatment showed a lesser development of earthworms, higher bacterial and fungal biomass

than vermicomposting treatment and greater differences, compared with the control without

earthworms, in cellulase, β-glucosidase, alkaline and acid phosphatase. Both treatments were

suitable for the stabilization of municipal sewage sludge and the combined composting-

vermicomposting treatment can be a viable process for maturation of fresh compost.

Keywords: enzyme activities, PLFAs, microbial community, earthworm, organic waste,

stability

Capítulo 3

91

1. INTRODUCTION

Municipal wastewater treatment plants produce significant amounts of sewage sludge,

amount to about 11 million dry tonnes per year in the EU, which needs suitable and

environmentally accepted management before final disposal (Kelessidis and Stasinakis,

2012). Sewage sludge can harm the environment when it is deposited directly on soil due to

its fermentative capacity and the presence of hazardous substances, both organic and

inorganic, including pathogenic organisms and heavy metals (Williams, 2005). Due to its high

organic load, sewage sludge is a suitable waste for being treated by biological techniques

such as composting and vermicomposting aimed at obtaining a stable product with a high

agronomic value.

Vermicomposting is a bio-oxidation and stabilization process of organic matter as a

result of the interaction between microorganisms and earthworms. Microorganisms are

mainly responsible for the degradation of organic material, although earthworms stimulate

microorganisms due to the modification of substrate properties through feeding, aeration

and cast excretion, which leads to the acceleration of mineralization of organic matter and the

improvement of nutrient availability for plants (Domínguez, 2004). Vermicomposting has

been successfully applied on the treatment of municipal sewage sludge. Most research on

sewage sludge vermicomposting has focused on the study of physical-chemical parameters

such as nutrients (Domínguez and Gómez-Brandón, 2013; Fu et al., 2015), humic and fulvic

substances (Zhang et al., 2015) and heavy metals (Suthar, 2010). Nevertheless, less is

reported on the microbiological and biochemical changes that occur during the

vermicomposting of municipal sewage sludge. Benitez et al. (1999) observed a reduction in β-

glucosidase, protease, urease and dehydrogenase activities related to the decline in available

substrates, in the first 6 weeks of vermicomposting of municipal sewage sludge mixed with

paper mill sewage sludge. Domínguez and Gómez-Brandón (2013) found that the presence of

the earthworm Eisenia andrei increased microbial biomass, measured as N-microbial biomass

by fumigation-extraction method, from week 1 to week 16 of vermicomposting of sewage

sludge compared to the control without earthworms. On the contrary, Fu et al. (2015)

observed a decrease of C-microbial biomass in the first 40 days of vermicomposting of

pelletized dewatered sludge, with subsequent low and constant values that indicated the

stability of the final products.

The integration of composting and vermicomposting has been considered a suitable

method for waste management (Ndegwa and Thompson, 2001). The inoculation of

earthworms in the material that passed through the thermophilic phase of composting has

Capítulo 3

92

been used as a pre-treatment before vermicomposting, in order to remove compounds

harmful for earthworms, such as ammonium (Domínguez, 2004). Several authors have

investigated the combined use of composting and vermicomposting for the treatment of

different organic materials, showing that prior composting can accelerate degradation and

improve the stabilization of the final product (Frederickson et al., 1997; Lazcano et al., 2008).

Fornes et al. (2012) studied the evolution of composting, vermicomposting and the combined

composting-vermicomposting process of horticulture waste, focusing their research on the

physical-chemical changes over time. These authors showed that vermicomposts had better

properties as growing media than as compost. Also, Lazcano et al. (2008) compared the

products of these processes, but not the evolution over time, by the study of microbiological

and biochemical parameters, noting that the combined process was the most effective

method for the stabilization of the cattle manure. Hait and Tare (2011) reported that the

combined composting-vermicomposting process of sewage sludge made it possible to obtain

a good quality pathogen-free product. Likewise, Sen and Chandra (2009) showed that

earthworms changed the dynamic of the bacterial community in the combined process

compared with composting of sugar waste, but they did not study the vermicomposting

process. So, no research on microbiological evolution over time of the vermicomposting

process and the combined composting-vermicomposting process for the same waste was

found.

It has been observed that the study of enzyme activities is a reliable index of the

evolution of organic matter during vermicomposting (Benitez et al., 1999; Aira et al., 2007a).

Enzyme activities provide information on the conversion of complex organic compounds into

more readily assimilable substances and, hence, enzymes are of interest to evaluate

stabilization throughout waste biodegradation. Likewise, enzyme activities have been related

to earthworm growth. Thus, they have been proposed as indicators to optimize

vermicomposting process (Fernández-Gómez et al., 2010). Benitez et al. (1999)

demonstrated that hydrolytic enzyme activities tended towards stability in the course of

sewage sludge vermicomposting, but however, the lack of controls makes it difficult to

distinguish between the effect caused by earthworms and the effect of the microbiota present

in the waste.

Conversely, phospholipid fatty acid analysis (PLFAs) is a useful tool for monitoring the

microbial community dynamics. The total amount of PLFAs can be used as an indicator of

viable microbial biomass and some PLFAs are specific to certain living organisms and,

therefore, can be used as biomarkers for the presence and abundance of microbial groups

Capítulo 3

93

(Zelles, 1999). Thus, analysing PLFAs during vermicomposting makes it possible to know the

changes in the microbial community composition over time. Fernández-Gómez et al. (2013)

observed a reduction in total PLFAs of different organic wastes after 24 weeks of

vermicomposting. In the same way, Gómez-Brandón et al. (2011a, 2013) reported that the

activity of earthworms reduced the PLFAs characteristic of bacterial and fungal biomass. This

reduction was more pronounced between week 21 and week 36 of rabbit manure

vermicomposting and pig slurry vermicomposting.

In this work, we studied the microbiological evolution during vermicomposting

compared with vermicomposting after composting, for the treatment and stabilization of

sewage sludge. The main hypothesis was that earthworms cause a different effect on

microbial community structure, depending on whether they feed on fresh or previously

composted material. To this end, enzyme activities (cellulase, β-glucosidase, protease,

alkaline and acid phosphatase) and the structure of the microbial community by analysing

PLFAs were assessed throughout the vermicomposting of fresh and composted sewage

sludge with the earthworm specie Eisenia andrei. In order to discern the effects due to the

different microbiological composition of the waste from the effects caused by earthworms,

the same substrates incubated without earthworms were studied.

2. MATERIALS AND METHODS

2.1. Substrates and earthworms

Sewage sludge was collected from a municipal wastewater treatment plant in Cangas

(Pontevedra, NW Spain) after an aerobic biological treatment and subsequent dehydration.

The sludge was mixed with wood chips as a bulking agent, adjusting the ratio to 1:2 (v/v). A

part of this mixture was used for the vermicomposting treatment (V). Another part of the

mixture was subjected to composting in a static adiabatic reactor with a 600 L capacity and

automatic control of temperature and oxygen. Forced aeration was applied, using a

centrifugal fan intermittently and depending on the controlled variables. The temperature

was maintained above 45°C for 7 days with maximum values of 60°C. The process ended after

15 days when the temperature in the composting mass reached values below 35°C. The fresh

compost was removed from the reactor, mixed and used as a substrate for vermicomposting

in the combined composting-vermicomposting treatment (CV). The earthworm species E.

andrei (Bouché, 1972) was used for the vermicomposting due to its high tolerance to

environmental factors and its high rate of organic matter processing (Domínguez, 2004). In

order to determine if the substrates affected the growth and maturation of the earthworms,

Capítulo 3

94

juvenile specimens with an average weight of 310 ± 25 mg were collected from a laboratory

culture fed with horse manure.

2.2. Experimental design

Vermicomposting was carried out in rectangular culture systems of 14 L capacity,

which were filled with a layer of sieved and moistened vermiculite as refuge for earthworms,

with the advantage of being a biologically inert material. A plastic mesh (5 cm mesh size) was

placed between the vermiculite and the substrate to prevent their mixture and facilitate the

sampling. Two kilograms of substrate sludge or compost (2000 ± 6 g) and 115-120

earthworms according to the feed rate of 0.75 kg feed/kg worm/day (Ndegwa et al., 2000),

were introduced. Each substrate was replicated three times. Controls involving the same

materials (vermiculite, mesh and sludge or compost) incubated without earthworms were

included in triplicate. Culture systems were kept in darkness under the same conditions. The

moisture content was controlled and maintained above 70% by watering throughout the

process. After 70 days, cocoons, earthworms and hatchlings were removed by hand from the

cultures, counted and weighed. The culture systems were maintained until day 112 to enable

the maturation of the vermicompost. Samples were taken at 0, 14, 28, 42, 56, 70, 91 and 112

days. In order to remove the bulking agent, samples were sieved (less than 10 mm) and

several parameters were determined, as detailed below.

2.3. Physical-chemical analysis

Organic matter content was measured by the loss on ignition of dried samples at 550ºC

for 4 hours. Inorganic nitrogen (N-NH4+ and N-NO3−) was determined in 0.5 M K2SO4 extracts

in a ratio of 1:10 (w/v) applying the modified indophenol blue colorimetric method (Sims et

al., 1995). Total extractable nitrogen was determined in the same extracts after oxidation

with K2S2O8, as described by Cabrera and Beare (1993), and dissolved organic nitrogen

content (DON) was calculated as (total extractable N) – (inorganic N). Total nitrogen content

(TN) and total carbon content (TC) were determined by combustion of dried samples using a

LECO 2000 CN elemental analyser. Water soluble carbon content (WSC) was analysed in

aqueous extracts 1:5 (w/v) after oxidation with H2SO4 (96%) and K2Cr2O7 (1N) at 160ºC for

30 minutes and spectrophotometric measurement of reduced chromium. The pH was

determined in aqueous extracts 1:10 (w/v) using a pH meter Crison Basic 20.

Capítulo 3

95

2.4. Biological and biochemical analysis

β-glucosidase was estimated by incubating the sample (1 g fresh weight) with 1 mL of

p-nitrophenyl-β-D-glucopiranoside (0.025 M) for 1 h at 37ºC and subsequent colorimetric

measurement of p-nitrophenol released (Eivazi and Tabatabai, 1988). Alkaline and acid

phosphatase was measured by incubating the sample (0.5 g fresh weight) with 1 mL of p-

nitrophenylphosphate (0.015 M) for 1 h at 37ºC and subsequent colorimetric measurement

of p-nitrophenol released (Eivazi and Tabatabai, 1977). Protease activity was measured by

colorimetric determination of the amino acids released, after the incubation of the sample (1

g fresh weight) with 5 mL of sodium caseinate (2%) for 2 h at 50ºC, using Folin-Ciocalteu

reagent (Ladd and Butler, 1972). Cellulase activity was assessed by colorimetric

determination of reducing sugars released after incubation of the sample (5 g fresh weight)

with 15 mL of carboxymethyl cellulose sodium salt (0.7%) for 24 h at 50ºC (Schinner and Von

Mersi, 1990). Germination index (GI) was calculated according to Zucconi et al. (1981) by

determining seed germination and root length of Lepidium sativum growing in 2 mL of

aqueous extracts 1:5 (w/v) in Petri dishes lined with paper filter during 48 hours.

The microbial community composition and biomass was determined by phospholipid

fatty acid analysis (PLFAs) following the method described by Gómez-Brandón et al. (2010)

for organic samples. Briefly, total lipids were extracted by stirring from 200 mg of each

freeze-dried sample with 60 mL of chloroform–methanol (2:1, v/v) and separated into

neutral lipids, glycolipids and phospholipids on silicic acid columns. The phospholipid

fraction was subjected to derivatization with trimethylsulfonium hidroxyde (TMSH) and fatty

acid methyl esters (FAMEs) obtained were analysed by gas chromatography and mass

spectrometry (GC-MS). GC-MS analysis was performed on a column CP-Select FAME, 100 m x

0.25 mm. FAMEs were identified by comparison of their retention time and mass spectra with

known standards (Larodan Fine Chemicals AB, Malmo, Sweden). The quantification was

performed using internal standard calibration. PLFAs were used to estimate the biomass of

specific microbial groups: gram-positive bacteria (i14:0, i15:0, a15:0, i16:0, a17:0), gram-

negative bacteria (16:1ω7, cy17:0, 17:1ω7, 18:1ω7, cy19:0) and fungi (18:2ω6, 18:1ω9,

20:1ω9)(Frostegård and Bååth, 1996; Zelles, 1997; Madan et al., 2002). The total amount of

PLFAs identified (totPLFAs) was used as an indicator of the viable microbial biomass (Zelles,

1999).

Capítulo 3

96

2.5. Statistical analysis

Microbiological and biochemical data were analysed by repeated measures analysis of

variance (ANOVAR) in which the type of material and the presence or absence of earthworms

were set as between-subjects factors, and time was set as within-subjects factor. Correlation

analyses were carried out to examine the relationships between PLFAs and enzyme activities.

Student's t-tests were performed to determine the difference between the two types of

materials used in vermicomposting and one-way analysis of variance (ANOVA) to determine

the difference between physical-chemical parameters at the end of the vermicomposting

process. For post hoc comparison between groups in the case of a significant effect, HSD

Tukey tests were used. Where the assumptions of normality and variance homogeneity were

not met, data were log-transformed. All statistical tests were evaluated at the 95% confidence

level using the SPSS 20.0 program.

3. RESULTS AND DISCUSSION

3.1. Growth and reproduction of E. andrei

As shown in Table 1, the survival of E. andrei in both treatments V and CV was high (>

96.5% in all culture systems) and no significant differences were found (p > 0.05), so fresh

and composted sewage sludge presented good properties for their management by

vermicomposting.

Table 1. Growth, sexual development and survival of E. andrei after 70 days of vermicomposting

(V) and composting-vermicomposting (CV). Values are mean ± standard error (n = 3).

V CV

Earthworm biomass (g) 54.8 ± 1.2a 45.3 ± 0.6b

Weight gain per earthworm (mg) 169.4 ± 5.1a 80.7 ± 6.2b

Matured earthworms (%) 95.0 ± 1.3a 86.3 ± 2.3b

No. of hatchlings 517 ± 54a 10 ± 1b

No. of cocoons 840 ± 42a 282 ± 11b

Survival (%) 97.7 ± 1.1a 99.4 ± 0.3a

Means with the same letter are not significantly different (paired-

sample Student’s t-test, p < 0.05).

Capítulo 3

97

The growth of E. andrei in V treatment was significantly larger than in VC treatment (t =

9.063, p < 0.05) and earthworm biomass significantly increased relative to the beginning of

the process, about 51.3% in V (t = 16.206, p < 0.01) and 25.6% in CV (t = 14.042, p < 0.01).

The highest number of mature earthworms, cocoons and hatchlings were obtained in V

treatment, with significant differences between treatments (t = 4.706, p < 0.05; t = 9.347, p <

0.05; t = 16.818, p < 0.05, respectively). These results suggested that the vermicomposting of

sewage sludge showed slightly better conditions for the development of E. andrei than

composted sludge. Frederickson et al. (1997) found that the growth rate of the epigeic

earthworm Eisenia fetida was reduced in pre-composted green waste compared with

vermicomposting of fresh material, suggesting that the nutritional content was more rapidly

decreased during the early stages of composting. Gunadi and Edwards (2003) propounded

that pre-composting of cattle manure can reduce bioavailable nutrients for earthworms,

inhibiting the growth rate and the number of cocoons and hatchlings produced by E. fetida.

Similar results were established in this study for municipal wastewater sewage sludge.

Moreover, CV treatment presented greater concentrations of harmful compounds for

earthworms, such as ammonium (Table 2), which could negatively affect earthworm

development.

3.2. Physical-chemical parameters

The physical-chemical properties at the initial and final materials are shown in Table 2.

The organic matter content was significantly lower in the presence of earthworms than in

controls (F1,11 = 62.483, p < 0.0001). After vermicomposting, organic matter decreased

about 14.6% in V treatment with earthworms and 3.4% in treatment without earthworms,

while reductions were about 12.7% and 6.6% in CV treatment with and without earthworms,

respectively. The presence of earthworms accelerated the mineralization of organic matter

(Elvira et al., 1996) so that the highest earthworm biomass in V treatment may have

produced a greater decrease in this measure. A greater reduction was noted in organic matter

in VC controls compared to V controls, which showed a higher loss of organic matter through

microbial breaking down after the passage of sewage sludge at the thermophilic phase of

composting. Ryckeboer et al. (2003) pointed out that the taxonomic and metabolic diversity

of bacteria and available substrates for fungi increase after the thermophilic phase of

composting.

Capítulo 3

98

Table 2. Initial and final physical-chemical properties of vermicomposting (V) and composting-

vermicomposting (CV). Values are means ± standard error (n = 3).

Time V_E.andrei V_control CV_E.andrei CV_control

OM (g) 0 495.0 ± 1.1a 507.4 ± 0.5b

112 421.2 ± 1.0a 480.2 ± 0.7b 443.5 ± 1.1c 473.1 ± 4.1b

pH 0 6.2 ± 0.1a 7.1 ± 0.0b

112 5.7 ± 0.0a 5.6 ± 0.0a 5.9 ± 0.0b 6.0 ± 0.1b

C to N ratio 0 10.2 ± 0.2a 9.6 ± 0.1a

112 9.9 ± 0.3a 10.5 ± 0.3ab 10.5 ± 0.3ab 11.2 ± 0.2b

TC (g kg-1dw) 0 428.0 ± 1.9a 401.8 ± 1.6b

112 319.6 ± 7.1a 350.1± 7.6b 323.4 ± 5.5a 349.2 ± 5.1b

TN (g kg-1dw) 0 39.0 ± 0.6a 38.9 ± 0.3a

112 32.3 ± 0.2ab 33.4 ± 0.5a 30.7 ± 0.4b 31.2 ± 1.0ab

WSC (g kg-1dw) 0 2.64 ± 0.13a 7.53 ± 0.33b

112 2.88 ± 0.11a 4.43 ± 0.10b 3.02 ± 0.10a 3.42 ± 0.05c

DON (g kg-1dw) 0 4.57 ± 0.08a 5.16 ± 0.15b

112 6.56 ± 0.08a 7.88 ± 0.22b 3.84 ± 0.23c 3.95 ± 0.13c

NH4+ (g kg-1dw) 0 0.69 ± 0.04a 2.41 ± 0.14b

112 0.20 ± 0.01a 0.37 ± 0.03b 0.28 ± 0.02c 0.37 ± 0.02b

GI (%) 0 65.8 ± 2.3a 33.3 ± 1.1b

112 92.1 ± 0.4a 76.7 ± 0.9b 97.9 ± 0.4c 80.2 ± 0.5d

OM: organic matter, dw: dry weight, TC: total carbon, TN: total nitrogen, WSC: water soluble carbon,

DON: dissolved organic nitrogen, GI: germination index. Means with different letter in the same row

are significantly different (Tukey HSD, p < 0.05).

All final products had acidic conditions and there were no significant differences

between the presence and absence of earthworms, but V treatment showed lower pH than CV

(F1,11 = 24.666, p < 0.0001). These results agreed with those obtained by other authors

(Ndegwa et al., 2000; Khwairakpam and Bhargava, 2009; Hait and Tare, 2011), who observed

a decrease in pH after vermicomposting of sewage sludge due to the formation of organic

acidic compounds and the mineralization of nitrogen and phosphorus. The decrease in

carbon and nitrogen content, due to the degradation and mineralization of organic matter by

microorganisms, earthworms and the joint action of both, maintained the C to N ratio at a low

level in all treatments, with similar initial and final values. This ratio is extensively used as an

indicator of maturity of organic waste, although Yadav et al. (2010) reported that the C to N

ratio should not be used as a maturity parameter for the vermicomposting if the original

Capítulo 3

99

waste is rich in nitrogen. A significant reduction of TN and TC was observed between initial

and final values with significant differences between all treatments (F3,11 = 153.409, p <

0.0001). With respect to TN, it was reduced by 21% on average in CV and 16% on average in

V, finding significant differences between V and CV treatments (F3,11 = 10.935, p < 0.01). The

low C to N ratio in the substrates and the aeration provided by the bulking agent may cause

the release of ammonia gas and decrease of TN in vermicomposting (Benitez et al., 1999;

Domínguez, 2004). In the case of DON, significant differences between V and CV (F1,11 =

98.193, p <= 0.0001) were detected with a reduction in CV and an increase in V in both

treatments with earthworms and controls. The presence of E. andrei showed lower

concentrations of ammonium (F1,11 = 0.437, p < 0.01), WSC (F1,11 = 16.131, p < 0.01) and CT

(F1,11 = 23.478, p < 0.01) than controls. These results were consistent with the general

hypothesis that earthworms promote carbon and nitrogen mineralization (Domínguez,

2004). Although readily assimilable carbon and nitrogen content increased after composting,

the subsequent process of vermicomposting diminished these parameters, favouring the

stabilization of combined CV treatment. Also, the presence of E. andrei showed significantly

higher GI than controls (F1,11 = 21.023, p < 0.01) and the CV treatment presented the greatest

value (97.9%). According to Zucconi et al. (1985), values of germination index greater than

80% present no phytotoxicity, so that both treatments with earthworms were effective for

eliminating phytotoxic substances, suggesting a suitable level of maturation. The physical-

chemical parameters evaluated and the germination index showed that earthworms

improved the sewage sludge properties, reaching optimal values of stabilization and

maturation in both V and CV treatments.

3.3. Enzyme activities

In general, the five hydrolytic enzymes studied decreased throughout all treatments

(Fig. 1, Fig. 2) accordingly, with metabolic degradation processes of organic matter

diminishing during vermicomposting. Benitez et al. (1999) observed similar trends during

vermicomposting of sewage sludge, suggesting that the decrease in hydrolytic activities

indicated stabilization of organic matter.

Cellulase and β-glucosidase are enzymes of the carbon cycle that play an important role

in the breakdown of organic matter. Cellulases degrade cellulose, releasing reducing sugars

and β-glucosidases catalyze the hydrolysis of β-glycosidic bonds of the carbohydrates (Alef

and Nannipieri, 1995). The cellulase activity decreased after composting and significant

differences were observed between V and CV, both at the outset (t = 4.455, p < 0.05) and

Capítulo 3

100

during vermicomposting (F1,8 = 8.607, p < 0.05) (Fig. 1a and 1b). Owing to cellulose being

broken down by thermophilic organisms (Ryckeboer et al., 2003), the decrease in substrates

after composting may have reduced cellulase activity. In V treatment, with and without

earthworms, cellulase activity diminished in a similar way over the time, whereas in CV

treatment the decline was clearly higher in the presence of earthworms compared to controls

from day 28. Earthworms significantly reduced cellulase activity (F1,8 = 40.530, p < 0.0001)

producing a significant interaction between treatment, time and earthworms

presence/absence (F6,48 = 4.302, p < 0.01). Therefore, not only the presence of E. andrei

resulted in a decrease of cellulase activity, but also the effect caused by time and the type of

substrate, that probably affected the available substrate for this enzyme and cellulolytic

microbial community. Due to the fact that earthworms can feed on fungi (Schönholzer et al.,

1999), the main consumers of cellulose, changes in fungal community by earthworms, as has

been observed in this study by the analysis of PLFAs, could affect the production of cellulases.

Likewise, Gómez-Brandón et al. (2011b) found that earthworm activity reduced cellulase

enzyme after vermicomposting of grape marc, suggesting that the vermicomposted material

reached a high degree of stabilization.

Fig. 1. Changes in cellulase and β-glucosidase activities with and without E. andrei in

vermicomposting (V) (a, c) and composting-vermicomposting (CV) (b, d). Values are mean ±

standard error (n = 3).

Capítulo 3

101

Regarding β-glucosidase, no significant differences between V and CV were detected at

the beginning of vermicomposting (t = -0.251, p > 0.05), these being about 4200 µg PNP g-1dw

h-1 (Fig. 1c and 1d). β-glucosidase activity presented significant differences between V and CV

during vermicomposting (F1,8 = 36.319, p < 0.0001), producing a significant interaction

between treatment, time and earthworms presence/absence (F6,48 = 5.374, p < 0.0001). There

was a gradual decline in enzyme β-glucosidase in V treatment with a minimum at the end of

the experiment (85% reduction), both in the presence and absence of earthworms. In the CV

treatment, the decline in the first sampling involved about 60% reduction of the enzyme

activity, with and without earthworms, and then the activity remained stable from 56 days.

Similar results were obtained by Benitez et al. (1999) during vermicomposting of sewage

sludge, showing a sharp decrease in β-glucosidase activity during the first 6 weeks with a

subsequent stabilization trend as a result of the decrease in available organic substrates.

Moreover, when the availability of C and N required for enzyme synthesis is limiting,

microorganisms can constrain their production of enzymes (Allison and Vitousek, 2005).

Sewage sludge showed a low C to N ratio (Table 2), which may lead to carbon becoming a

limiting substrate for the microbiota. These results showed differences between treatments

for enzymes of the C cycle, indicating that enzyme activities decreased and stabilized before

in combined composting-vermicomposting than in vermicomposting of fresh sewage sludge.

Alkaline and acid phosphatase enzymes catalyze the hydrolysis of organic

phosphomonoester to inorganic phosphorus, differing according to their optimum pH of

activity. At the beginning of vermicomposting, significant differences were observed between

V and CV for both acid phosphatase (t = 7045, p < 0.05) (Fig. 2a and 2b) and for alkaline (t =

3442, p < 0.05) (Fig. 2c and 2d). Differences in the activities were also detected between the

two treatments for the two enzymes throughout the process (F1,8 = 7.147, p < 0.05 for acid

phosphatase; F1,8 = 6.153, p < 0.05 for alkaline phosphatase). Phosphatase activities

decreased with time, but because phosphatases are highly influenced by the pH (Eivazi and

Tabatabai, 1977), a further decrease of the alkaline phosphatase with values greater than

60% reduction was observed. The accumulation of inorganic compounds of phosphorus as a

result of enzymatic activity has been shown to repress phosphatase activity (Alef and

Nannipieri, 1995). Likewise, Aira et al. (2007a) suggested that a decrease in microbial

biomass may cause an increase of available phosphorus and, therefore, a decrease in

phosphatase activity. This drop in phosphatase activity was also observed by other authors in

the vermicomposting of different wastes (Fernández-Gómez et al., 2010, 2013). In V

treatment (Fig. 2a and 2c), phosphatase activities evolved in a similar way, regardless of the

presence or absence of earthworms. However, a greater influence of E. andrei on enzyme

Capítulo 3

102

activity was observed in the CV treatment (Fig. 2b and 2d) with a faster decrease in

phosphatase activities in the presence of earthworms, resulting in significant interaction

between treatment, time and earthworms presence/absence in acid (F6,48 = 10.810, p <

0.0001) and alkaline phosphatase (F6,48 = 1.961, p <0.05). Furthermore, earthworms

significantly affected alkaline phosphatase (F1,8 = 24.109, p < 0.01). Earthworms had a greater

effect on the enzymes of the phosphorus cycle in the combined CV treatment than V

treatment. Microbiota decrease and mineralization increase during composting, with

subsequent activity of earthworms, may have increased the content of orthophosphate and

reduced the activity and synthesis of phosphatase in CV treatment.

Fig. 2. Changes in acid phosphatase, alkaline phosphatase and protease activities with and

without E. andrei in vermicomposting (V) (a, c, e) and composting-vermicomposting (CV) (b, d,

f). Values are mean ± standard error (n = 3).

Capítulo 3

103

Protease enzymes supply a large part of the available nitrogen by catalyzing the

hydrolysis of proteins to peptides and amino acids (Alef and Nannipieri, 1995; Geisseler and

Horwath, 2008). Significant differences between V and CV (t = 11.046, p < 0.01) were

detected at the beginning of vermicomposting (Fig. 2e and 2f). Protease activity decreased in

all treatments with a greater reduction in V compared to CV (F1,8 = 36.986, p < 0.0001 ).

Protease activity is dependent on substrate availability (Aira et al., 2007a), which was

reduced as the vermicomposting process developed. Also, the greater activity in the final

samples of CV treatment could be caused by the formation of complexes between

extracellular enzymes and humic substances because of the previous process of composting

that can increase the formation of these substances compared to vermicomposting

(Campitelli and Ceppi, 2008). The evolution through time in V treatment was similar with and

without earthworms, showing a decrease in the first 14 days of process, remaining stable up

to 70 days and decreasing around 5000 µg tyrosine g-1dw 2h-1 in the final sampling. In CV

treatment without earthworms, protease activity slightly increased over time until it reduced

from 70 days to similar values to those observed at the beginning of the process. In the case

of CV treatment with earthworms, two different stages were observed: a first stage where

activity increased to a maximum at 28 days (17700 µg tyrosine g-1dw 2h-1) and a second

stage, from 42 days to 112 days, where the activity fell below the values noted at the

beginning of the process. No significant differences between treatments with earthworms

and controls were detected, although there was a significant interaction between treatment,

time and earthworms presence/absence (F6,48 = 9.273, p < 0.0001). In general, a reduction in

protease was observed over time, more marked in the treatment V and with a slight negative

effect of earthworms on this enzymatic activity. Several studies have observed reductions of

protease activity in the presence of earthworms in contrast to controls with different organic

substrates (Aira et al., 2007a; Gómez-Brandón et al., 2011b; Fernández-Gómez et al., 2013).

Geisseler and Horwath (2008) suggested that microorganisms regulate protease synthesis,

depending on their needs of carbon and nitrogen, so that the low C to N ratio could cause a

decrease in microbial biomass and enzyme synthesis.

3.4. Dynamic of the microbial community

Microbial biomass presented significant differences between V and CV (t = 21.719, p <

0.01) at the beginning of vermicomposting with a totPLFAs content in the first case of 1280

µg g-1dw, and after the composting process this was reduced up to 895 µg g-1dw. Both

treatments showed a decrease in microbial biomass throughout the vermicomposting

Capítulo 3

104

process (Fig. 3), especially in the first samplings, resulting in a significant interaction between

treatment and time (F6,48 = 70.315, p < 0.0001).

Fig. 3. Changes in microbial biomass, measured as totPLFAs, with and without E. andrei in

vermicomposting (V) and composting-vermicomposting (CV). Bars represent standard errors.

Different letters in the same time of sampling are significantly different (Tukey post hoc test p <

0.05).

This decline in microbial biomass, measured as totPLFAs, was consistent with the

results obtained by other authors during vermicomposting (Fernández-Gómez et al., 2013;

Gómez-Brandón et al., 2013). The sewage sludge used in this experiment was biologically

degraded during processing at the plant, so that the easily degradable nutrients could

exhaust, thereby resulting in a reduction of microbial biomass. Earthworm activity greatly

reduced the abundance of totPLFAs (F1,8 = 117.789, p < 0.0001), although this effect was

more pronounced in V treatment (98.5% reduction from initial) compared with CV (89.4%

reduction from initial), producing a significant interaction between treatment and presence

/absence of earthworms (F1,8 = 6.469, p < 0.05). Because V treatment presented a higher

increase in earthworm population and biomass compared to that observed in CV treatment,

microbial biomass showed a larger decline in totPLFAs. These findings are consistent with

previous observations that suggest that the digestion of organic material by epigeic

earthworms has negative effects on microbial biomass (Gómez-Brandón et al., 2011c). In

addition, Tiunov and Scheu (2004) observed that earthworms can compete with

microorganisms for available carbon resources and cause a sharp decline in microbial

biomass when carbon is limiting. The reduction of available resources due to the action of

Capítulo 3

105

microorganisms, competition for them with earthworms and feeding with bacteria and fungi

by earthworms, may explain the higher decline in biomass in the presence of E. andrei. In V

treatment with earthworms, totPLFAs correlated positively with all enzymes: cellulase (r =

0.787, p <0.0001), alkaline phosphatase (r = 0.832, p < 0.0001), β-glucosidase (r = 0.942, p <

0.0001), acid phosphatase (r = 0.684, p < 0.001) and protease (r = 0.750, p < 0.0001), with a

slightly lower correlation in all enzymes (p < 0.01) in absence of earthworms. In CV

treatment with earthworms, correlation between totPLFAs with cellulase (r = 0.791, p <

0.0001), glucosidase (r = 0.649, p < 0.01), protease (r = 0.481, p < 0.05), alkaline (r = 0.822, p

< 0.0001) and acid phosphatase (r = 0.601, p < 0.001) were observed. No correlations were

found, however, between totPLFAs and acid phosphatase and protease in CV control, whereas

correlations were detected with cellulase (r = 0.533, p < 0.01), alkaline phosphatase (r =

0.763, p < 0.01) and β- glucosidase (r = 642, p < 0.01). Despite the different evolution of

hydrolytic enzymes, the results in general showed that a decrease in microbial biomass was

accompanied by a decrease in enzymes, suggesting that enzyme activities were directly

associated with living microorganisms. However, V treatment with E. andrei showed enzyme

activities very similar to controls, despite having a stronger reduction of microbial biomass.

This may indicate the formation of complexes between enzymes and humic substances

during vermicomposting (Benitez et al., 2005) or the presence of a less efficient microbial

community for organic matter degradation in controls (Aira et al., 2007b). Although

earthworms exerted an important effect on the microbial biomass, the performance of the

enzyme activity was determined by both the type of waste used in the vermicomposting and

the presence or absence of earthworms. Both treatments were suitable for the reduction and

stabilization of microbial biomass of sewage sludge.

Table 3. Changes in PLFAs without earthworms (control) and in the presence of E. andrei in vermicomposting (V) and composting-vermicomposting

(CV). Values are means ± standard error (n = 3).

Time Gram+ bacteria (µg g-1dw ) Gram– bacteria (µg g-1dw ) Fungi (µg g-1dw )

V CV V CV V CV

E.andrei control E.andrei control E.andrei control E.andrei control E.andrei control E.andrei control

0 197.0±8.8a 178.5±7.2a 370.6±14.9a 270.1±16.8b 338.2±15.6a 186±17.2b

14 274.9±12.1a 270.9±16.9a 89.2±10.6b 160.2±5.4c 313.0±14.4a 321.3±13.7a 113.7±11.3b 184.6±13.0c 214.7±11.1a 217.8±15.3a 62.5±8.1b 105.8±10.8c

28 121.3±7.9a 159.2±10.1b 42.6±3.5c 97.7±4.6a 134.2±7.7a 211.0±13.1b 38.4±3.8c 102.3±9.8a 111.1±9.1a 159.9±13.0b 34.6±3.5c 78.4±10.2d

42 38.0±4.8a 98.5±2.5b 28.5±1.9a 59.1±4.4c 28.4±4.7a 144.7±10.2b 26.6±4.0a 57.8±5.2c 34.5±5.1a 115.5±6.5b 26.4±3.6a 39.0±4.6a

56 42.5±1.6a 96.4±9.5b 27.3±4.3c 44.0±5.6ac 41.6±1.2a 107.1±1.9b 14.3±2.7c 52.0±5.0a 46.4±1.8a 95.0±2.6b 19.4±2.5c 38.7±4.0a

70 27.9±2.4a 77.5±2.2b 23.4±2.6a 53.3±3.1c 18.7±3.1a 54.1±4.9b 13.5±3.0a 51.3±2.8b 22.5±4.9ac 69.2±3.1b 17.0±5.4c 43.2±7.0a

91 11.4±1.6a 38.5±4.0b 17.6±3.6a 27.8±0.7b 8.9±1.5a 25.9±2.2b 10.5±1.7a 24.0±0.6b 10.3±1.4a 47.1±3.5b 9.6±1.9a 26.4±0.4c

112 2.8±0.1a 52.4±5.9b 12.6±2.1c 16.8±1.0c 0.4±0.1a 33.3±3.2b 11.0±1.2c 26.5±2.2b 0.9±0.3a 69.8±8.6b 8.1±1.4c 36.6±2.2d

In each parameter different letters in the same time of sampling are significantly different (Tukey post hoc test p < 0.05).

Capítulo 3

107

Earthworm activity had a strong effect on the abundance of PLFAs with a significant

reduction in Gram + bacteria (F1,8 = 134.119, p < 0.0001), Gram – bacteria (F1,8 = 106.862, p <

0.0001) and fungi (F1,8 = 76.260, p < 0.0001) with regard to control during vermicomposting

(Table 3). Similar results have been reported by Gómez-Brandón et al. (2011a) during

vermicomposting of pig slurry with E. fetida, suggesting that this earthworm modified the

structure of microbial communities. In the same way, E. andrei affected the abundance of

bacterial and fungal PLFAs during vermicomposting of fresh and composted sewage sludge.

Previous studies have observed that epigeic earthworms had a greater effect on Gram +

bacteria than Gram – bacteria through the gut associated processes and that Gram – bacteria

can survive the transit through the earthworm gut (Gómez-Brandón et al., 2011c; Williams et

al., 2006). In accordance with this, V treatment with earthworms showed a higher reduction

in Gram + bacteria than Gram – bacteria compared to controls throughout the process.

Likewise, E. andrei had a greater effect on Gram + bacteria in the first samplings of CV

treatment. Also, a significant decrease in the fungal biomass was observed in the presence of

earthworms. This effect on fungal populations was reported by other authors. Huang et al.

(2013) and Fernández-Gómez et al. (2013) found a decrease in fungal biomass in the end

products after vermicomposting with E. fetida. Such decreases may be due to the fact that

earthworms can feed on fungi that provide them an important source of nutrients

(Schönholzer et al., 1999). The decrease of microbial groups was greater in V treatment with

earthworms (> 98.6%) than in CV treatment with earthworms (93% Gram + and 96% for

Gram – and fungi), while the reduction in controls was higher in CV treatment for Gram +

bacteria (90.6%) compared to V treatment (73.4%) and similar for Gram – bacteria (90-91%)

and fungi (79-80%), showing a significant interaction between treatment, time and

earthworms presence/absence for all indicators of specific microbial groups: Gram + (F6,48 =

8.697, p < 0.01), Gram – (F6,48 = 6.216, p < 0.01) and fungi (F6,48 = 4,989, p < 0.01). Regarding

the last samplings, it was observed that treatments with earthworms, both in V and CV, had a

lower concentration of bioindicators of microbial groups than the controls. It has been found

that the degradation and mineralization of organic matter in the controls of vermicomposting

is insufficient compared to treatment with earthworms (Elvira et al., 1996). So, the greatest

microbial load in controls may indicate the presence of organic matter available for the

microorganisms, being lower in CV treatment due to its passage through the thermophilic

phase of composting.

As can be seen in the last sampling, final products presented different microbial

communities. Lores et al. (2006) reported that the fingerprint of the microbial community of

a vermicompost depends on the type of substrate and earthworm species used in the process.

Capítulo 3

108

Fernández-Gómez et al. (2012) reported that vermicomposts produced from wastes of

different nature and origin can contain similar microbial communities, if the same earthworm

species is used. Sen and Chandra (2009) showed divergent bacterial community in compost

and vermicompost obtained from the same initial waste, despite similar changes in their

physico-chemical parameters during the processes. In this case, the microbial community

seemed to differ depending on if V or CV treatment were applied and on the presence or

absence of earthworms. Differences may be attributable to the distinct physicochemical and

microbiological composition of the waste and the interactions of earthworms on microbial

communities.

4. CONCLUSIONS

In this study, the ability of the earthworm Eisenia andrei to degrade fresh and

composted municipal sewage sludge was shown. The PLFA analysis indicated that E. andrei

reduced microbial biomass, both bacterial and fungal, and that the composition of the

microbial community depended on the presence or absence of earthworms as well as the

treatment used. Earthworm incorporation after the thermophilic stage of composting

accelerated the degradation of the waste, with a greater decrease in the phosphorus and

carbon cycling enzymes. Therefore, the combined composting-vermicomposting process can

be a good alternative to the management of sewage sludge.

ACKNOWLEDGMENTS

This study was financially supported by the Xunta de Galicia (Regional Autonomous

Government of Galicia) (09MDS024310PR). The authors thank the wastewater treatment

plant of Cangas for supplying sewage sludge and the research support services of the

University of Vigo (CACTI) for the carbon and nitrogen analysis. The authors also thank

Emilio Rodríguez Cochón for technical support and referees for their valuable comments and

suggestions.

Capítulo 3

109

REFERENCES

Aira, M., Monroy, F., Domínguez, J., 2007a. Earthworms strongly modify microbial biomass and activity triggering enzymatic activities during vermicomposting independently of the application rates of pig slurry. Sci. Total Environ. 385, 252-261.

Aira, M., Monroy, F., Domínguez, J., 2007b. Eisenia fetida (Oligochaeta: Lumbricidae) modifies the structure and physiological capabilities of microbial communities improving carbon mineralization during vermicomposting of pig manure. Microb. Ecol. 54, 662–671.

Alef, K., Nannipieri, P., 1995. Methods in applied soil microbiology and biochemistry. Academic Press, London.

Allison, S.D., Vitousek, P.M., 2005. Responses of extracellular enzymes to simple and complex nutrient inputs. Soil Biol. Biochem. 37, 937-944.

Benitez, E., Nogales, R., Elvira, C., Masciandaro, G., Ceccanti, B., 1999. Enzyme activities as indicators of the stabilization of sewage sludges composting with Eisenia foetida. Bioresour. Technol. 67, 297-303.

Benitez, E., Sainz, H., Nogales, R., 2005. Hydrolytic enzyme activities of extracted humic substances during the vermicomposting of a lignocellulosic olive waste. Bioresour. Technol. 96, 785-790.

Cabrera, M.L., Beare, M.H., 1993. Alkaline persulfate oxidation for determining total nitrogen in microbial biomass extracts. Soil Sci. Soc. Am. J. 57, 1007–1012.

Campitelli, P., Ceppi, S., 2008. Effects of composting technologies on the chemicals and physicochemical properties of humic acids. Geoderma 144, 325–333.

Domínguez, J., 2004. State of the art and new perspectives in vermicomposting research, in: Edwards, C.A. (Ed.), Earthworm Ecology, second ed. CRC Press, Boca Raton, pp. 401–424.

Domínguez, J., Gómez-Brandón, M., 2013. The influence of earthworms on nutrient dynamics during the process of vermicomposting. Waste Manage. Res. 31, 859-868.

Eivazi, F., Tabatabai, M.A., 1977. Phosphatases in soils. Soil Biol. Biochem. 9, 167–172.

Eivazi, F., Tabatabai, M.A., 1988. Glucosidases and galactosidases in soils. Soil Biol. Biochem. 20, 601–606.

Elvira, C., Goicoechea, M., Sampedro, L., Mato, S., Nogales, R., 1996. Bioconversion of solid paper-pulp mill sludge by earthworms. Bioresour. Technol. 57, 173-177.

Fernández-Gómez, M.J., Nogales, R., Insam, H., Romero, E., Goberna, M., 2010. Continuous-feeding vermicomposting as a recycling management method to revalue tomato-fruit wastes from greenhouse crops. Waste Manage. 30, 2461–2468.

Fernández-Gómez, M.J., Nogales, R., Insam, H., Romero, E., Goberna, M., 2012. Use of DGGE and COMPOCHIP for investigating bacterial communities of various vermicomposts produced from different wastes under dissimilar conditions. Sci. Total Environ. 414, 664-671

Fernández-Gómez, M.J., Díaz-Raviña, M., Romero, E., Nogales, R., 2013. Recycling of environmentally problematic plant wastes generated from greenhouse tomato crops through vermicomposting. Int. J. Environ. Sci. Technol. 10, 697-708.

Fornes, F., Mendoza-Hernández, D., García-de-la-Fuente, R., Abad, M., Belda, R.M., 2012. Composting versus vermicomposting: a comparative study of organic matter evolution through straight and combined processes. Bioresour. Technol. 118, 296–305.

Capítulo 3

110

Frederickson, J., Butt, K.R., Morris, R.M., Daniel, C., 1997. Combining vermiculture with traditional green waste composting systems. Soil Biol. Biochem. 29, 725–730.

Frostegård, Å., Bååth, E., 1996. The use of phospholipid fatty acid analysis to estimate bacterial and fungal biomass in soil. Biol. Fertil. Soils 22, 59-65.

Fu, X., Huang, K., Chen, X., Li, F., Cui, G., 2015. Feasibility of vermistabilization for fresh pelletized dewatered sludge with earthworms Bimastus parvus. Bioresour. Technol. 175, 646-650.

Geisseler, D., Horwath, W.R., 2008. Regulation of extracellular protease activity in soil in response to different sources and concentrations of nitrogen and carbon. Soil Biol. Biochem. 40, 3040-3048.

Gómez-Brandón, M., Lores, M., Domínguez, J., 2010. A new combination of extraction and derivatization methods that reduces the complexity and preparation time in determining phospholipid fatty acids in solid environmental samples. Bioresour. Technol. 101, 1348–1354.

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011a. Changes in microbial community structure and function during vermicomposting of pig slurry. Bioresour. Technol. 102, 4171–4178.

Gómez-Brandón, M., Lazcano, C., Lores, M., Domínguez, J., 2011b. Short-term stabilization of grape marc through earthworms. J. Hazard. Mater. 187, 291-295.

Gómez-Brandón, M., Aira, M., Lores, M., and Domínguez, J. 2011c. Epigeic earthworms exert a bottleneck effect on microbial communities through gut associated processes. PLoS One 6, 1-9.

Gómez-Brandón, M., Lores, M., Domínguez, J., 2013. Changes in chemical and microbiological properties of rabbit manure in a continuous-feeding vermicomposting system. Bioresour. Technol. 128, 310-316.

Gunadi, B., Edwards, C.A., 2003. The effects of multiple applications of different organic wastes on the growth, fecundity and survival of Eisenia fetida (Savigny) (Lumbricidae). Pedobiol. 47, 321-329.

Hait, S., Tare, V., 2011. Vermistabilization of primary sewage sludge. Bioresour. Technol. 102, 2812-2820.

Huang, K., Li, F.S., Wei, Y.F., Fu, X.Y., Chen, X.M., 2013. Changes of bacterial and fungal community compositions during vermicomposting of vegetable wastes. Bioresour. Technol. 150, 235–241.

Kelessidis, A., Stasinakis, A.S., 2012. Comparative study of the methods used for treatment and final disposal of sewage sludge in European countries. Waste Manage. 32, 1186-1195.

Khwairakpam, M., Bhargava, R., 2009. Vermitechnology for sewage sludge recycling. J. Hazard. Mater. 161, 948-954.

Ladd, J.N., Butler, J.H.A., 1972. Short-term assays of soil proteolytic enzyme activities using proteins and dipeptide derivatives as substrates. Soil Biol. Biochem. 4, 19–30.

Lazcano, C., Gómez-Brandón, M., Domínguez, J., 2008. Comparison of the effectiveness of composting and vermicomposting for the biological stabilization of cattle manure. Chemosphere 72, 1013-1019.

Lores, M., Gómez-Brandón, M., Pérez-Díaz, D., Domínguez, J., 2006. Using FAME profiles for the characterization of animal wastes and vermicomposts. Soil Biol. Biochem. 38, 2993-2996.

Madan, R., Pankhurst, C., Hawke, B., Smith, S., 2002. Use of fatty acids for the identification of AM fungi and the estimation of the biomass of AM spores. Soil Biol. Biochem. 34, 125-128.

Ndegwa, P.M., Thompson, S.A., Das, K.C., 2000. Effects of stocking density and feeding rate on vermicomposting of biosolids. Bioresour. Technol. 71, 5-12.

Capítulo 3

111

Ndegwa, P.M., Thompson, S.A., 2001. Integrating composting and vermicomposting in the treatment and bioconversion of biosolids. Bioresour. Technol. 76, 107-112.

Ryckeboer, J., Mergaert, J., Vaes, K., Klammer, S., De Clercq, D., Coosemans, J., Insam, H., Swings, J., 2003. A survey of bacteria and fungi occurring during composting and self-heating processes. Ann. Microbiol. 53, 349–410.

Schinner, F., Von Mersi, W., 1990. Xylanase-, CM-cellulase- and invertase activity in soil: an improved method. Soil Biol. Biochem. 22, 511–515.

Schönholzer, F., Hahn, D., Zeyer, J., 1999. Origins and fate of fungi and bacteria in the gut of Lumbricus terrestris L. Studied by image analysis. FEMS Microbiol. Ecol. 28, 235–248.

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.

Sims, G.K., Ellsworth, T.R., Mulvaney, R.L., 1995. Microscale determination of inorganic nitrogen in water and soil extracts. Commun. Soil Sci. Plant Anal. 26, 303–316.

Suthar, S., 2010. Pilot-scale vermireactors for sewage sludge stabilization and metal remediation process: comparison with small-scale vermireactors. Ecol. Eng. 36, 703-712.

Tiunov, A.V., Scheu, S., 2004. Carbon availability controls the growth of detritivores (Lumbricidae) and their effect on nitrogen mineralization. Oecologia 138, 83-90.

Williams, P.T., 2005. Waste treatment and disposal, second ed. John Wiley & Sons, Chichester.

Williams A.P., Roberts P., Avery L.M., Killham K., Jones D.L., 2006. Earthworms as vectors of Escherichia coli O157:H7 in soil and vermicomposts. FEMS Microbiol. Ecol. 58, 54-64.

Yadav, K.D., Tare, V., Ahammed, M.M., 2010. Vermicomposting of source-separated human faeces for nutrient recycling. Waste Manage. 30, 50–56.

Zelles, L., 1997. Phospholipid fatty acid profiles in selected members for soil microbial communities. Chemosphere 35, 275-294.

Zelles, L., 1999. Fatty acid patterns of phospholipids and lipopolysaccharides in the characterisation of microbial communities in soil: a review. Biol. Fertil. Soils 29, 111–129.

Zhang, J., Lv, B., Xing, M., Yang, J., 2015. Tracking the composition and transformation of humic and fulvic acids during vermicomposting of sewage sludge by elemental analysis and fluorescence excitation–emission matrix. Waste Manage. 39, 111-118.

Zucconi, F., Pera, A., Forte, M., de Bertoldi, M., 1981. Evaluating toxicity of immature compost. BioCycle 22, 54-57.

Zucconi, F.; Monaco, A.; Forte, M.; De Bertoldi, M. 1985. Phytotoxins during the stabilization of organic matter. In: Composting of agricultural and other wastes, pp. 73-5. Gasser J.K.R. (ed.). Elsevier, London.

CAPÍTULO 4

Product quality and microbial dynamics during

vermicomposting and maturation of compost from pig

manure

Revista: Waste Management

Estado: bajo revisión

Permiso: pre print

Capítulo 4

115

RESUMEN

El estiércol de cerdo es un residuo ganadero que se produce en elevadas cantidades y

que presenta una alta carga orgánica pudiendo ser tratado mediante procesos biológicos. El

principal objetivo de este trabajo es estudiar, a través de la dinámica microbiana, el proceso

de vermicompostaje y el empleo de lombrices de tierra en la maduración del compost fresco,

y la influencia de estos procesos en la calidad de los compost y vermicompost. Para

determinar el efecto que las lombrices de tierra causan en la dinámica microbiana, se

evaluaron las actividades enzimáticas y los perfiles de ácidos grasos fosfolípidos a lo largo de

112 días. Así mismo se analizaron parámetros fisicoquímicos y biológicos en los productos

obtenidos. La presencia de lombrices de tierra redujo significativamente la biomasa

microbiana y todos los grupos microbianos (bacterias Gram +, bacterias Gram – y hongos)

tanto en el estiércol fresco como en el pre-compostado. Las actividades enzimáticas (celulasa,

β-glucosidasa, fosfatasa ácida y proteasa) se comportaron de manera significativamente

distinta dependiendo del tratamiento efectuado. La comunidad microbiana tuvo un efecto

directo sobre los procesos degradativos durante la maduración del compost, sin embargo, en

presencia de Eisenia andrei se establecieron complejas interacciones lombrices-microbiota-

sustrato. Así, las actividades enzimáticas en presencia de lombrices no fueron consecuencia

directa de la síntesis de enzimas por parte de la microbiota presente en el sustrato. La

inoculación de lombrices de tierra en el compost mejoró la calidad del producto obtenido por

lo que el proceso combinado compostaje-vermicompostaje es un proceso viable para la

maduración del compost. Aunque el vermicompostaje del estiércol fresco alcanza valores de

calidad, es necesario optimizar el tiempo de la fase de maduración del vermicompost.

Palabras clave: actividades hidrolíticas, PLFAs, evolución microbiana, madurez, curado

del compost, lombrices de tierra.

Capítulo 4

117

ABSTRACT

Pig manure is a livestock waste product which is produced in large quantities and has a

high organic load, which can be treated through biological processes. The main objective of

this research is to study, through microbial dynamics, the vermicomposting process and the

use of earthworms in the maturation of fresh compost, and the influence of these processes

on compost and vermicompost quality. In order to determine the effect that earthworms have

on microbial dynamics, the enzymatic activities and profiles of phospholipid fatty acids

(PLFAs) were evaluated over a 112-day period. The physicochemical and biological

parameters of the obtained products were also analysed. The presence of earthworms

significantly reduced microbial biomass and all the microbial groups (Gram + bacteria, Gram

– bacteria, and fungi), both in fresh and the pre-composted manure. The enzymatic activities

(cellulase, β- glucosidase, acid phosphatise, and protease) behaved in a significantly

distinctive manner, depending on the treatment used. The microbial community had a direct

effect on the degradative processes during compost maturation. However, complex

earthworm-microbiota-substrate interactions were established in the presence of Eisenia

andrei. Likewise, the enzymatic activities in the presence of earthworms were not a direct

consequence of in situ enzyme synthesis by the microbiota present in the substrate. The

inoculation of the earthworms in the compost improved the quality of the product obtained,

which means that the combined process of composting-vermicomposting is a viable process

for maturing compost manure. Although the vermicomposting of fresh manure achieved

quality values, it is necessary to optimize the vermicompost maturation phase period in

terms of time.

Keywords: hydrolytic activities, PLFAs, microbial evolution, mature, compost curing,

earthworms

Capítulo 4

119

1. INTRODUCTION

Spain is the second largest pork producer in the European Union and generates more

than 50 million m3 of pig slurry per year as a result of this activity; this then must be

correctly dealt with (MAGRAMA, 2012). The separation of the slurry’s liquid phase from its

solid one is a treatment process regularly carried out on pig farms (Hjorth et al., 2010). The

solid portion is mainly composed of excrement, straw, and the remains of food with a high

content of organic matter, which allows for it to be treated through biological techniques,

such as vermicomposting and composting. Vermicomposting is a process which bio-oxidizes

and stabilizes the organic material due to the interaction between microorganisms and

earthworms. The microorganisms are the ones principally responsible for degrading the

organic material, although the earthworms stimulate the microorganisms as a result of them

modifying the substrate’s properties through feeding, aeration, and the excretion of casts; all

of this leads to the acceleration of the mineralization of the organic material and an

improvement in the nutrient availability for plants (Domínguez, 2004). The vermicomposting

process can generally be separated into two phases. First, an active phase, characterized by

the earthworms which move through the waste, causing changes to its physiochemical and

microbial properties. This is followed by the second phase, the maturation stage, when the

earthworms move towards fresher material; this results in the microorganisms taking

control of decomposing the processed material (Lazcano et al., 2008). Various authors have

studied the effects of epigeic earthworms have on pig manure. For example, Aira et al. (2007)

studied the interaction between the earthworms and microbial biomass and enzymatic

activity by using continuous feed vermireactors, with different doses of manure in the

presence and absence of Eisenia fetida. In a later study Aira and Domínguez (2009) compared

microbial stabilization and nutrients of the casts produced by E. fetida fed with pig manure

with those fed with cow manure. Similarly, Gómez-Brandón et al. (2011a) studied the impact

the species E. andrei has on the structure and function of the microbial community in the

casts from three types of animal manure, among them pig manure. Monroy et al. (2009)

studied the effect of the E. fetida earthworm on the total number of coliforms in pig manure in

continuous feed vermireactors, observing a reduction of 98% of coliforms density.

Composting is a process of biological degradation of solid organic substrates under

aerobic conditions through the action of different microbial communities, resulting in a stable

and humified product and apt for adding to the soil (Insam and de Bertoldi, 2007). The

organic material goes through different stages: a mesophilic stage, characterized by

microbiota proliferation; a thermophilic stage, which produces a high level of biodegradation,

Capítulo 4

120

the growth of thermophilic organisms and the inhibition of non-thermo-tolerant organisms;

and the final stage, which includes a period of cooling, stabilization, and maturation, and is

characterized by the growth of mesophilic organisms and the humification of the compost

(Ryckeboer et al., 2003b). The inoculation of earthworms in the material after going through

the thermophilic stage has been used as a pre-treatment prior to vermicomposting, with the

aim being to eliminate compounds present in some waste products which are noxious for

earthworm; these include, for example, high levels of ammonia (Domínguez, 2004). For

example, (Chan and Griffiths (1988) faced with the impossibility of vermicomposting the pig

manure without treating it, so prepared it through composting and the addition of calcium

sulphate. The comparison of vermicomposting with regards to composting pig manure was

dealt with by Zhu et al. (2016) who compared both treatments with the objective of studying

the variation of heavy metals and dissolved organic matter. No studies about the inoculation

of earthworms in pig manure after a previous composting with the aim of knowing about the

changes during maturation, with or without epigeic earthworms, and comparing the compost

and vermicompost obtained have been found.

This paper studies the microbial dynamics during the vermicomposting of pig manure

and during the maturation process of pig manure compost, with and without epigeic E. andrei

earthworms. The main hypothesis of this study are 1) that the presence of detritivores

earthworms determines the changes to the microbiota and thus the changes in the enzymatic

activity and that 2) maturing the compost with earthworms improves the product obtained in

terms of physic-chemistry and biology. For this, enzymatic activity (cellulase, β-glucosidase,

protease, and acid phosphatase) and the microbial community’s structure were studied by

analyzing PLFAs throughout the vermicomposting and compost maturation process. The

stability and maturity parameters in the products obtained were also analyzed.

2. MATERIALS AND METHODS

2.1. Substrates and earthworms

The solid fraction of manure was collected from a pig breeding farm after storage in

manure pits for liquid separation. The manure was mixed with shredded wood as a bulking

agent, adjusting the ratio to 1:2 (v/v). The initial properties of this mixture can be seen in

table 1 as fresh pig manure. One part was used for the vermicomposting treatment with

earthworms (V treatment) and the control treatment without earthworms (Vcontrol treatment).

Another part was subjected to composting in a static adiabatic reactor with a 600 L capacity

and automatic control that has been described in detail in Villar et al. (2016a). The

Capítulo 4

121

temperature was maintained above 45°C for 8 days with maximum values of 60°C. The

process ended after 14 days when the temperature in the composting mass reached values

below 35°C. The fresh compost was removed from the reactor, mixed and one part was used

as a substrate for vermicomposting in the combined composting-vermicomposting treatment

(CV treatment) and the other part continued the composting maturation process without

earthworms (C treatment). The initial properties of the pre-composted pig manure can be

seen in the table 1.

Table 1. Physico-chemical characteristics of the substrates used in the treatments

Fresh pig manure

Pre-composted pig

manure

Moisture (%) 76.2 ± 0.5 68.8 ± 0.3*

Organic matter (%) 87.3 ± 0.2 85.7 ± 0.8*

pH 8.40 ± 0.03 6.86 ± 0.03*

Electrical conductivity (mS cm-1) 0.86 ± 0.02 0.64 ± 0.02*

Total carbon (mg g-1dw) 433.1 ± 11.2 419.0 ± 6.9*

Total nitrogen (mg g-1dw) 23.2 ± 0.4 22.5 ± 1.3

P2O5 (mg g-1dw) 11.3 ± 0.3 11.7± 0.4

Ammoniacal nitrogen (mg g-1dw) 3.02 ± 0.10 1.40 ± 0.05*

The asterisk indicates significant differences between substrates (paired-sample

Student’s t-test, p < 0.05) dw: dry weight

Specimens with an average weight of 325 ± 34 mg of the earthworm species E. andrei

(Bouché, 1972) were collected from a laboratory culture fed with horse manure for the CV

and V treatments.

2.2. Experimental design

Vermicomposting treatments were carried out according with previous work (Villar et

al., 2016b). Briefly, rectangular culture systems of 14 L capacity were filled with a layer of

sieved and moistened vermiculite, a plastic mesh of 5 cm mesh size and 2010 ± 8 g of

substrate: pig manure or compost. Each substrate was replicated three times and 115-120

earthworms were introduced in each system. Controls involving the vermiculite, mesh and

fresh pig manure incubated without earthworms were included in triplicate for Vcontrol

treatment. Culture systems were kept in darkness under the same conditions. The moisture

content was controlled and maintained above 70% by watering throughout the process. After

Capítulo 4

122

70 days, cocoons, earthworms and hatchlings were removed by hand from the cultures,

counted and weighed. The culture systems were maintained until day 112 to enable the

maturation of the vermicompost. Maturation of the pre-composted pig manure without

earthworms (C treatment) was done in triplicate in wooden boxes of 200L as has been

described in Villar et al. (2016a). The compost was left for 112 days without being

homogenized and it was not necessary to moisten the material. Therefore, four different

treatments were carried out (Table 2). Samples were taken at 0, 14, 28, 42, 56, 70, 91 and 112

days in composting boxes and vermicomposting systems. In order to remove the bulking

agent, samples were sieved (less than 10 mm) and several parameters were determined, as

detailed in the following sections.

Table 2. Abbreviations, substrates and presence/absence of earthworms from the experimental

design

Treatments Substrate Earthworms

V Fresh pig manure yes

Vcontrol Fresh pig manure no

CV Pre-composted pig manure yes

C Pre-composted pig manure no

2.3. Physicochemical analysis

Organic matter content was measured by the loss on ignition of dried samples at 550 º

C for 4 hours. Total carbon content (TC) and total nitrogen content (TN) were determined by

combustion of dried samples using a LECO 2000 CN elemental analyser. N-NH4+ and N-NO3−

were determined in 0.5 M K2SO4 extracts in a ratio of 1:10 (w/v) according with Sims et al.

(1995). Water soluble carbon content (WSC) was analysed in aqueous extracts 1:5 (w/v) by

dichromate oxidation in sulfuric acid solution and spectrophotometric measurement of

reduced chromium. The pH and electrical conductivity were determined in aqueous extracts

1:10 (w/v) using a pH meter Crison Basic 20 and a conductivimeter Crison CM 35.

2.4. Biological and biochemical analysis

β-glucosidase and acid phosphatase were determined by the colorimetric measurement

of p-nitrophenol released according to the methods reported by (Eivazi and Tabatabai, 1988,

1977). Protease activity was measured by colorimetric determination of the amino acids

Capítulo 4

123

released using Folin-Ciocalteu reagent (Ladd and Butler, 1972). Cellulase activity was

assessed by colorimetric determination of reducing sugars released according with Schinner

and von Mersi (1990). Germination index (GI) was calculated according to Zucconi et al.

(1981) by determining seed germination and root length of Lepidium sativum growing in

aqueous extracts 1:5 (w/v).

Static respiration rate (SR) was measured using manometric respirometers by

OxiTop® system (WTW GmbH, Weilheim, Germany). The microbial community composition

and biomass were determined by phospholipid fatty acid analysis (PLFAs) following the

method described by Gómez-Brandón et al. (2010) for organic samples. The analysis was

performed with a CP-Select FAME, 100m x 0.25mm in a gas chromatograph-mass

spectrometer (GC-MS). Identification was done by comparison of retention times and mass

spectra with known external standards (Larodan Fine Chemicals AB, Malmo, Sweden) and

quantification was performed with methyl nonadecanoate fatty acid (C19:0) as internal

standard. PLFAs were used to estimate the biomass of specific microbial groups: Gram-

positive bacteria (i14:0, i15:0, a15:0, i16:0, a17:0), Gram-negative bacteria (16:1ω7, 17:1ω7,

18:1ω7) and fungi (18:2ω6, 18:1ω9, 20:1ω9). The total amount of PLFAs identified

(totPLFAs) was used as an indicator of the viable microbial biomass (Zelles, 1999).

2.5. Statistical analysis

All statistical tests were performed using R software (R Development Core Team,

2014). Student's t-tests were performed to determine the difference between earthworm-

related parameters and substrates and one-way analysis of variance were performed to

determine difference between final products. Enzyme and PLFA data was analyzed with

linear mixed-effects models using the nlme package (Pinheiro et al., 2015). The type of

substrate and the presence or absence of earthworms were fixed factors and the repeated

measurement throughout time in each maturation box was treated as a random effect to

address the non-independence of samples. Logarithmic and square root transformations of

the data were necessary to ensure the normality and homogeneity of the variance of residuals

of models. For post hoc comparison between treatments, Tukey tests were carried out using

the glht function of the multcomp package (Hothorn et al., 2008). Correlation analyses were

also carried out to examine the relationships between PLFAs and enzyme activities with cor

function of the stats package. All statistical tests were evaluated at the 95% confidence level

and values are given as the mean ± standard deviation.

Capítulo 4

124

3. RESULTS

3.1. Growth and reproduction of E. andrei

Table 3 shows the parameters related to the growth, reproduction, and sexual

development of E. andrei in fresh manure and compost after leaving the reactor. Both mixes

make it possible for high level of earthworms to survive, although there are significant

differences (p < 0.05) due to the higher mortality rate in the earthworm population in the

treatment with compost as the substrate. Earthworm growth, the number of mature

individuals and cocoon production were similar in both substrates, there being no significant

difference between the two (p > 0.05). With respect to the number of juveniles, the V

treatment showed significantly more individuals than the CV treatment (p < 0.05).

Table 3. Growth, sexual development and survival of E. andrei after 70 days of

vermicomposting.

V treatment CV treatment

Earthworm biomass (g) 40.8 ± 0.3 37.0 ± 2.4*

Weight gain per earthworm (mg) 44.6 ± 2.4 43.1 ± 6.9

Mature earthworms (%) 96.8 ± 1.2 94.7± 1.1

No. of hatchlings 1208 ± 102 955 ± 43*

No. of cocoons 548 ± 67 463 ± 59

Survival (%) 98.8 ± 1.3 85.8 ± 3.7*

The asterisk indicates significant differences between substrates (Student’s t-test, p <

0.05)

3.2. Microbial dynamics evolution

3.2.1. Microbial activity

The substrates showed significant differences at 0 time with respect to the enzymes:

greater cellulose (p< 0.0001) and β-glucosidase (p< 0.05) activity and less acid phosphatase

(p < 0.0001) and protease (p < 0.0001) activity in the compost as compared to the fresh

manure. All the enzymatic activity studied (β-glucosidase, cellulase, acid phosphatase, and

protease) showed significant differences (p < 0.01), depending on the treatment; this resulted

in significant interactions between the treatment and time (p <0.0001) (Figure 1).

Capítulo 4

125

Cellulase activity was significantly affected by the presence of earthworms (p < 0.001)

this effect being dependent on the substrate (p < 0.05). The presence of earthworms in fresh

manure caused a decrease in cellulase activity, as compared to the Vcontrol treatment, which

had high cellulase activity during the first 91 days of the process. The presence of E. andrei

caused an initial increase in cellulase activity in the CV treatment up until day 28, but after

that saw a sharp decrease, resulting in values similar to the C treatment at the end of the

process. The C treatment had similar activity values throughout the entire maturation

process. After 112 days, the activities were C, CV > Vcontrol > V.

Fig. 1. Changes in cellulase (a), β-glucosidase (b), acid phosphatase (c) and protease (d)

activities during vermicomposting (V) and control (Vcontrol) treatments of pig manure,

maturation with earthworms (CV) and maturation without earthworms (C) of pre-composted

pig manure.

In terms of β-glucosidase activity, the presence/absence of earthworms had a

significant effect (p < 0.05) this effect being dependent on the substrate (p < 0.0001). In the

case of fresh manure, the inoculation of E. andrei caused a clear decrease in activity as

compared to the control without earthworms: the V treatment had slight fluctuations until

reaching values around 1000µg g-1 in the last samplings, while Vcontrol had large swings

throughout the process, but with values greater than V in every sample. The presence of

Capítulo 4

126

earthworms caused an initial decrease β-glucosidase activity in the CV treatment, and then

increased after the day-28 sampling, showing values which were clearly superior to C

treatment. Despite these differences, after 112 days V and C products showed values that

were similar to each other and less than those found in CV and Vcontrol, which also reached

similar values. No correlations between these two carbon-cycle enzymes were observed.

The presence/absence of earthworms significantly affected acid-phosphatase activity

(p < 0.001), independent of the type of substrate. The species E. andrei in both substrates

increased this enzyme’s activity to a greater degree than the treatments without earthworms.

At the end of the process, the activity in the Vcontrol and CV treatments were very similar, as

occurred with β-glucosidase, the values being higher than the activity in treatment V, while

treatment C showed activity in the last samplings of around 2000µg g-1, which are less than

the rest of the treatments.

The presence of earthworms in the substrates did not affect protease activity, although

significant interaction between the presence/absence and type of substrate was observed (p

< 0.01) as well as significant interaction between the presence/absence and time (p < 0.01).

Compost maturation with and without earthworms showed different evolutions after day 28,

greater activity being observed in CV than in C, while in the fresh manure the effects were to

the contrary and the presence of earthworms caused a decrease in protease activity over

time. In the last sampling, only the compost showed differences, with lower values; as in the

case of β-glucosidase and acid phosphatase, CV y Vcontrol products presented similar values.

The correlations were different according to the treatment, so that cellulase correlated

with protease in the V (r = 0.81, p < 0.0001), Vcontrol (r = 0.74, p < 0.0001), and CV (r = -0.52, p

< 0.001) treatments. Cellulase correlated with acid phosphatase in the V (r = -0.90, p <

0.0001), and Vcontrol (r = -0.81, p < 0.0001) treatments. Acid phosphatase negatively correlated

with protease in the V (r = -0.82, p < 0.0001), Vcontrol (r = -0.63, p < 0.01), and C (r = -0.87, p <

0.0001) treatments. β-glucosidase correlated with acid phosphatase in the CV (r = 0.72, p <

0.001) and C (r = -0.87, p < 0.001) treatments. β-glucosidase correlated with protease in the C

(r = 0.83, p < 0.0001) treatment.

3.2.2. Microbial communities.

The substrates showed significant differences at 0 time in all the assessed parameters;

totPLFAs (p < 0.0001), Gram + bacteria (p < 0.0001) and fungi (p < 0.0001) all had a

significant decrease in concentration after the composting process, while the Gram – bacteria

increased significantly (p< 0.05) after the manure went through the composting reactor. Both

Capítulo 4

127

the microbial biomass and the microbial groups studied over time showed significant

differences, depending on the treatment (p < 0.0001) resulting in significant interactions

between treatment and time (p <0.0001) (Figure 2).

Fig. 2. Changes in a) microbial biomass, b) Gram + bacteria, c) Gram – bacteria and d) fungi

estimated by phospholipid fatty acids (PLFA) analysis during vermicomposting (V) and control

(Vcontrol) treatments of pig manure, maturation with earthworms (CV) and maturation without

earthworms (C) of pre-composted pig manure.

TotPLFAs showed significant differences as a consequence of the presence or absence

of E. andrei (p < 0.001), independent of the type of substrate. The earthworms caused a

decrease in microbial biomass, although there was an increase after their removal at time 70.

In the case of the V treatment, concentration of totPLFAs increased by a factor of 2.8, whereas

the CV treatment saw an increase by a factor of 1.3. Both the Vcontrol and C treatments

maintained stable biomass levels after day 56, with a greater concentration in C (around 120

µg g-1) than in Vcontrol (around 100 µg g-1). So, it has been observed in the C treatment that

the totPLFAs is correlated with every enzyme: positively with β-glucosidase (r = 0.75, p <

0.0001) and protease (r = 0.75, p < 0.0001) and negatively with cellulase (r = -0.58, p < 0.01)

and acid phosphatase (r = -0.79, p < 0.0001). The enzymes in the V treatment did not

correlate with totPLFAs. In the Vcontrol treatment totPLFAs correlated with the protease (r =

Capítulo 4

128

0.56, p < 0.01) and the acid phosphatase (r = -0.48, p < 0.05) and in the CV treatment with the

β-glucosidase (r = -0.46, p < 0.0001) and cellulase (r = 0.57, p < 0.0001).

Likewise, the earthworms caused a decrease in the abundance of the PLFA biomarkers

for Gram + bacteria and, after their removal at time 70, both treatments gradually increased;

V treatment saw an increase of 4.4 times and CV treatment saw an increase of 2.3 times at the

end of the process. In this case, the Vcontrol treatment also increased the abundance of Gram +

bacteria in the final sampling, whereas the C treatment maintained stable levels, around 40

µg g-1, starting from the day 42 sampling. So, significant differences caused the

presence/absence of earthworms were observed (p < 0.01), independent of the type of

substrate. So, it has been observed that in the C treatment the Gram + bacteria correlated

with every enzyme: positively with the β-glucosidase (r = 0.76, p < 0.0001) and the protease

(r = 0.68, p < 0.001) and negatively with the cellulase (r = -0.56, p < 0.01) and the acid

phosphatase (r = -0.73, p < 0.001). The enzymes from the Vcontrol did not correlate with the

Gram + bacteria. The Gram + bacteria correlated with the β-glucosidase (r = 0.46, p < 0.05) in

the V treatment. In the CV treatment the PLFAs characteristic of Gram + bacteria correlated

with the cellulase (r = 0.50, p < 0.05) and the protease (r = -0.54, p < 0.05).

A reduction in the PLFAs characteristic of Gram – bacteria was observed, as a

consequence of the activity of E. andrei earthworms with respect to the treatments without

earthworms (p <0.001). This effect did not depend on the type of substrate, but rather that

the epigeic earthworms reduced to a great degree these PLFA biomarkers in the fresh

manure and after the removal of the earthworms the abundance increased until reaching

levels greater than the other treatments (increased of 3.8 times in V and 1.7 in CV). The

abundance of PLFAs characteristic of Gram – bacteria remained stable from day 56 in the

Vcontrol treatment. So, it has been observed that in the C treatment the Gram – bacteria

correlated with every enzyme: positively with the β-glucosidase (r = 0.81, p < 0.0001) and the

protease (r = 0.75, p < 0.0001) and negatively with the cellulase (r = -0.51, p < 0.05) and the

acid phosphatase (r = -0.75, p < 0.001). The enzymes in the V treatment did not correlate with

the Gram – bacteria. In the CV treatment they correlated with the cellulase (r = 0.50, p < 0.05)

and with the protease (r = -0.44, p < 0.05) in the Vcontrol treatment.

Just as in the previous parameters, significant differences of the abundance of

biomarkers characteristic of fungi in function of the presence/absence of E. andrei

earthworms were observed (p < 0.0001), independent of the type of substrate. The CV

treatment had a gradual decrease over time of the fungal biomass and the V treatment saw a

marked decrease until the removal of the earthworms on day 70, which produced a slight

Capítulo 4

129

increase (2.8 times). Both C and Vcontrol treatments saw decreases with greater fluctuation. So,

it has been observed that the C treatment the fungi correlated with every enzyme: positively

with the β-glucosidase (r = 0.63, p < 0.01) and the protease (r = 0.65, p < 0.001) and

negatively with the cellulase (r = -0.47, p < 0.05) and the acid phosphatase (r = -0.79, p <

0.0001). The PLFAs characteristic of fungi in the V treatment correlated with the cellulase (r =

0.47, p < 0.05), the protease (r = 0.58, p < 0.01) and the acid phosphatase (r = -0.63, p < 0.01).

In the Vcontrol the fungal biomass correlated with the protease (r = 0.55, p < 0.05), the cellulase

(r = 0.56, p < 0.01) and the acid phosphatase (r = -0.71, p < 0.001). In the CV treatment the

fungal biomass correlated with the β-glucosidase (r = -0.55, p < 0.01) and the acid

phosphatase (r = -0.80, p < 0.0001).

3.3. Compost parameters

With the exception of the C/N and WSC ratios, the stability and maturation parameter

results showed significant differences between treatments (p < 0.05) (Table 4).

The result of the individual PLFA cluster analysis (Figure 3) showed two groups clearly

formed by the vermicompost of the V treatment and another group formed by the products

from the other three treatments. In this last group there was the biggest difference between

the CV treatment and the Vcontrol and C treatments, there existing greater similarity between

products that had not been treated with earthworms.

Capítulo 4

130

Table 4. Parameters of stability and maturity of compost and vermicompost after 112 days of

process.

Fresh pig manure Pre-composted pig manure

Earthworms Without Earthworms Without

Organic Matter (%) 71.0 ± 0.6b 76.4± 1.3a 67.4 ± 1.6c 74.5 ± 1.5a

pH 7.06 ± 0.09a 6.51± 0.04c 6.82 ± 0.04b 6.28 ± 0.12d

EC(mS cm-1) 0.41 ± 0.1d 0.54 ± 0.0b 0.45 ± 0.0c 0.70 ± 0.12a

Total carbon (mg g-1dw) 351.2 ± 0.3d 407.5± 6.7a 364.9 ± 4.1c 390.3 ± 6.8b

Total nitrogen (mg g-1dw) 17.8 ± 0.5c 20.1± 1.9a 18.3 ± 0.3b 22.0 ± 1.2a

ratio C/N 19.7 ± 0.5 20.5± 2.2 19.9 ± 0.6 17.8 ± 1.2

WSC (mg g-1dw) 1.74 ± 0.06 2.00 ± 0.10 2.09 ± 0.09 2.21 ± 0.12

NH4+/NO3- 0.08 ± 0.01d 0.19 ± 0.01c 0.15 ± 0.02b 0.65 ± 0.03a

SR (mg O2 g-1OM h-1) 0.57 ± 0.01b 0.54 ± 0.03b 0.45 ± 0.02c 0.69 ± 0.03a

GI (%) 97.9 ± 3.5a 76.5 ± 3.4c 96.6± 3.3a 88.1 ± 9.1b

Parameters with different letters indicate significant differences between products (Tukey test, p <

0.05). Parameters without letters were not significantly different (ANOVA, p < 0.05). (EC: electrical

conductivity, WSC: water soluble carbon, SR: static respiration, GI: germination index).

Fig. 3. Dendrogram of cluster analysis based on PLFA profiles of the products obtained after 112

days of process. Vermicomposting (V) and control (Vcontrol) treatments of pig manure,

maturation with earthworms (CV) and maturation without earthworms (C) of pre-composted

pig manure.

Capítulo 4

131

4. DISCUSSION

4.1. Growth and reproduction of E. andrei

The results show the ability of the E. andrei earthworms to grow and reproduce both in

fresh pig manure and in pre-composted pig manure, which means that vermicomposting is

appropriate for treating pig manure and maturing after previously composting it.

From the point of view of the development and reproduction of E. andrei, maturing the

pre-composted manure with inoculated epigeic earthworms is a viable alternative to

maturing the compost sensu stricto. It has been noted that it is necessary or convenient to

carry out pre-composting for some waste products in order to avoid earthworm death by

passing the waste product through thermophilic temperatures in order to eliminate or

reduce compounds toxic to these detritivore organisms (Garg et al., 2005). In this case, the

pre-composting produced slightly noxious conditions in the substrate which caused a greater

death rate for the E. andrei specimens. Gunadi and Edwards (2003) proposed that the pre-

composting of cattle manure could reduce nutrients bioavailable to the earthworms,

inhibiting the growth and reproduction of E. fetida and Yadav et al. (2015) stated that the

quality and palatability of the food directly affected survival, the growth rate, and the

reproduction potential of the earthworms. Despite the fact that pre-composting reduces the

nutritional content for the earthworms, this reduction was not significant; the population

ended up developing adequately. (Villar et al., 2016b) showed that the presence of toxic

compounds in the pre-composted sewage sludge affected the growth and reproduction of E.

andrei. So, pre-composting created the inconvenience of generating compounds that were

toxic for E. andrei, although the recuperation and establishment of the earthworm population

indicated that these compounds were volatile and/or easily biodegradable, thus reducing

their toxicity within a short period of time.

4.2. Microbial dynamics evolution

The results showed distinct evolutions in the microbial biomass and microbial groups,

depending on the treatment used as reflected to a lesser or greater degree in the distinct

behavior of the enzymatic activity. The complex interactions that occurred between the

earthworm-microbiota-substrate had an effect on the degradative activity of the organic

compounds and the distinct microbial populations. The low rate of correlation observed

between the enzymatic activity and the different microbial groups verifies an important effect

of the synthesized and released enzymes on the environment throughout the process and

Capítulo 4

132

they remained active or inactive, in the medium or attached to colloids, depending on

environmental conditions and, to a lesser extent, as a consequence of enzymatic activity

synthesized in situ by the live microbiota present in the substrate. So, it is known that

enzyme behaviour depends on various factors that interact together as well as the location of

the enzymes which contribute to the activity (Nannipieri et al., 2002).

The hydrolytic activities presented a different behaviour between treatments, meaning

that the degradative activity was different in the vermicomposting and compost maturation,

depending on whether it was with or without earthworms.

The carbon-cycle enzymes studied are those related to the decomposition of cellulolytic

material: cellulase and β-glucosidase activity. The cellulase measurement jointly estimates

both endoclunase activity, which depolymerizes the cellulose, and the β-glucosidase, which

hydrolyzes cellobiose to make glucose. Likewise, β-glucosidase activity was determined

independently. The synthesis of these enzymes implies the induction by the substrate and

repression by an easily-used source of carbohydrates, such as glucose. Pig manure is a

material rich in this type of cellulolytic material, due the high levels of bedding and food

remains which form part of the manure. The earthworms cause a decrease in cellulase

activity both with vermicomposting of fresh manure as well as in compost maturation.

Gómez-Brandón et al. (2011c) observed a reduction in cellulase activity after incubating

grape marc with E. andrei, taking into account that this decrease could be due to the effect the

earthworms have on the microbial biomass, reducing enzyme synthesis, as the decrease of

available nitrogen and carbon compounds affecting enzymatic activity. This effect was

observed in the treatments of fresh manure such that a decrease in fungi cause a decline

cellulase enzyme production, there being greater in the presence of E. andrei, since the fungal

community is the main source of food for earthworms and these enzymes are largely

synthesized by fungi. Nonetheless, epigeic earthworms also affected the fungal community

during compost maturation, but a direct effect of the fungi on the cellulase was not observed.

It is true, in this case, that bacterial evolution, both Gram + and Gram –, affected the activity of

this enzyme. In this respect, Villar et al. (2016a) attributed cellulase activity during pig

manure compost maturation via turning system to the presence of a cellulolitic bacterial

community. In terms of compost maturing under static conditions, the maintenance of the

cellulase activity and its negative correlation with all microbial groups could indicate the

slow degradation of this polymer by a specialized microbial community and the presence of

extracellular free enzymes on the medium or joined to particles of the substrate. Inasmuch as

the variability and concentration of compounds decline during composting, it decreases the

Capítulo 4

133

microbiota; although its diversity may be greater and more effective at degrading the more

complex polymeric compounds (Ryckeboer et al., 2003b). When epigeic earthworms are

inoculated into pre-composted manure the bacteria take control of synthesizing this enzyme,

as already been mentioned, which shows that competition, interaction, and feeding of the

microorganisms on the part of the earthworms causes a selective filter on the microbiota and

that the remaining bacterial biomass is a community specialized in metabolizing cellulolytic

compounds. At the end of the process, the cellulase was greater in compost maturation

treatments; consequently, the passing of manure through the thermophilic stage meant a

short-term degradation of the most available materials, resulting in the concentration of

cellulolytic materials and cellulase synthesis.

The action of E. andrei earthworms on β-glucosidase activity had antagonistic effects,

depending on substrate that it fed from; the earthworms caused a reduction in activity in the

fresh manure, whereas the opposite was true for compost. The greater β-glucosidase activity

in the present of earthworms concurs with the results of other authors (Aira et al., 2007;

Benitez et al., 1999; Benítez et al., 2002; Vivas et al., 2009) who demonstrated an increase in

β-glucosidase activity, due to it being an enzyme inducible by substrate; the combined action

of microorganisms and earthworms can liberate low-molecular weight carbon molecules,

which increase this activity. According to this premise, β-glucosidase should have increased

as well during the vermicomposting of fresh manure, although the opposite effect was

observed. Benitez et al. (1999) and Villar et al. (2016b) showed a decrease in β-glucosidase

enzyme activity during sewage sludge, assuming that the reduction in activity was due to the

depletion of substrates and the low C/N ratio, which limited enzyme production. Nonetheless,

the increase in activity after removing the earthworms on day 70 suggests that the decrease

could be the consequence of the effect or combined interaction of: the consumption by the

earthworms of the substrates that could be hydrolyzed by this enzyme and, as a result, cause

a decline in the microbial population in the pig manure; and the direct consumption of the

microbiota by the earthworms. This interaction corresponds with the pronounced decline of

the microbial biomass and, in particular, of Gram + bacteria, which correlate with β-

glucosidase activity. Unlike that observed in the cellulase enzyme, the evolution of β-

glucosidase during the static maturation of compost was positively related to the evolution of

both the bacterial and fungal biomass. The lack of the influence of epeigeic earthworms on

the microbiota allowed for a direct relation between the synthesis of the β-glucosidase

enzyme by the microorganisms with their hydrolytic activity once compounds that were

easiest to degrade had been consumed in the prior process of composting.

Capítulo 4

134

The enzyme phosphatase is pH dependent and in all the treatments maintained pH

values which were close to neutral, but slightly acidic, despite the fresh manure initially

having an alkaline pH, which means that the increase in the acid phosphatase activity in all

the treatments is connected with the change in pH towards acidic conditions. Garcia et al.

(1992) related the increase of phosphatase activity with the high quantities of phosphorous

compounds present in some waste products. The pig manure used in this study showed total

phosphorous levels greater than the sewage sludge used by Garcia et al. (1992). What pigs

are fed contributes to the increase of the phosphorous contents in their excrement, mainly by

inorganic compounds (von Wandruszka, 2006). As such, the concentration of phosphate

compounds could induce the synthesis of this enzyme. Thus, the acidification of the medium

and the induction by substrate promoted phosphatase activity in every treatment, likewise

the increase in the mineralization of organic phosphorous also contributed to the decrease in

pH. Likewise, the presence of E. andrei caused phosphatase activity to be greater than in

substrates without earthworms. According to Satchell and Martin (1984) the increase of this

enzyme in casts may be due to the enzymes themselves excreted by the earthworms and

indirectly by stimulating the microbiota. Furthermore, Aira and Domínguez (2009) showed

that an increase in the inorganic phosphorous in casts could cause an increase in phosphatase

activity. The presence of E. andrei, both in manure as well as compost, reduced microbial

biomass; this indicates that phosphatase activity may be due to free enzymes, joined to

colloids or humic particles, or promoted by the earthworms, and that the activity persists in

conditions which inhibit the microbial biomass. Likewise, a negative correlation has been

found with the biomarkers for fungi in every treatment. Bacteria are more effective at

dissolving phosphorous than fungi, which means that the decline in the fungal biomass did

not negatively affect the phosphatase enzyme activity, reducing the competition in terms of

the types of phosphorous with the bacteria. Consequently, the smaller microbial biomass

maintained greater activity than the treatments without earthworms, suggesting that the

passage through the intestine could result in microbial population that, while smaller, is

metabolically more active (Gómez-Brandón et al., 2011b).

The enzyme protease is considered indicative of the state of degradation of organic

material due to its extreme dependence on the availability of substrate (Lazcano et al., 2008).

During compost maturation in static conditions both the correlation with the microbial

biomass and maintenance of protease activity over time suggest the stability of degradative

activities of the organic substrates directly acted upon by both the bacterial and fungal

microbiota present in the compost. Voběrková et al. (2017) observed greater protease

activity during fungi inoculation during composting, suggesting that proteolytic activity

Capítulo 4

135

seems to be strongly influenced by the microbial populations and that protein and

polypeptide hydrolysis increases during the maturation stage, which shows incomplete

protein decomposition during the biooxidative stage. It has been noted that the presence of

earthworms tends to reduce protease activity in various waste products (Aira et al., 2007;

Fernández Gómez et al., 2013; Gómez-Brandón et al., 2011b; Villar et al., 2016b). This effect

was clearly observed in the treatment of fresh manure with earthworms, while the

maturation of compost with E. andrei saw greater protease activity than in the compost

without earthworms. Parthasarathi and Ranganathan (2000) observed that protease activity

is greater in casts made by Eudrilus eugeniae and Lampito mauritii than in material that has

not been ingested. Ravindran et al. (2016) observed an increase in proteolytic bacteria in the

first weeks of vermicomposting; these then decreased afterwards as a result of a decline in

proteins. These results seem contradictory, since a greater number of protein compounds are

degraded during the thermophilic stage, decreasing the substrate for this enzyme in the pre-

composted manure. But after the removal of the earthworms, the protease activity in the

vermicompost slightly increased while that in the pre-composted manure decreased, which

could indicate a decline in protein compounds accompanied by a decline in proteolytic

bacteria and fungi.

In compost undergoing maturation, where the more assimilable compounds have

already been degraded in the thermophilic stage, both the bacterial and fungal microbial

biomass directly directs the degradation of the more recalcitrant organic material. The effect

of the earthworms makes it so hydrolytic activities are no directly driven by changes in the

microbial community of the substrate, especially in fresh manure.

The decrease in the microbial biomass, measured in totPLFAs, was consistent with the

results obtained by other authors during vermicomposting (Fernández Gómez et al., 2013;

Gómez-Brandón et al., 2013; Villar et al., 2016b) and composting (Garcia et al., 1992; Ros et

al., 2006; Villar et al., 2016a). The greater reduction of the microbial biomass with the

presence E. andrei is consistent with previous observations, which suggested that the

digestion of organic material by the earthworms has a negative effect on the microbial

communities (Gómez-Brandón et al., 2011a). The reduction in the available resources due to

microorganisms action, the competition for them with earthworms and the feeding of

bacteria and fungi by earthworms, may explain the greater reduction of biomass with the

presence of E. andrei (Villar et al., 2016b). This was demonstrated through removing the

competition and activity exerted by the earthworms on the microbial populations; the PLFA

biomarkers of the microbial biomass and groups increased, showing the presence of nutrition

Capítulo 4

136

available to the microbiota in the vermicomposting of fresh manure. In the maturing

compost, the microbial biomass was maintained from day 28. At this respect, Tiquia et al.,

(2002) found that the stabilization of ATP during composting may indicate the maturity of

the compost and suggest a change in the microbial community to more specialized microbial

groups and Villar et al., (2016a) suggested that the stability of the microbial biomass in the

turned maturation of pig manure could be due to the slow degradation of more recalcitrant

compounds. The stable conditions of the maturation process keep the microbial biomass

stable by slowing degrading the organic compounds present.

The increase of Gram + bacteria after removing the earthworms could suggest the

availability of organic substrates as a consequence of the breakage of lignocellulolytic

compounds (Elouaqoudi et al., 2015), the increase in compost maturation being less due to

the material going through the thermophilic stage that reduced nutritional content. Likewise

Gram – bacteria increased to a greater degree in the fresh manure following the removal of

the earthworms but also in pre-composted material and to a lesser extent in the static

compost. Kato et al. (2005) suggested that the proportion of Gram + PLFA markers could be

used as a tool to determine compost maturity. On the other hand, Danon et al. (2008) showed

that Bacteroidetes dominated at the start of compost maturation, followed by Actinobacteria,

and finally by Gammaproteobacteria. Neher et al. (2013) showed that Chloroflexi and c-

Proteobacteria decreased in relative abundance during maturation and were surpassed in

abundance by Bacteroidetes in the finished composts. Castillo et al. (2013) showed that, at the

end of the vermicompost maturation period, the relative abundance of Betaproteobacteria

and Actinobacteria experienced a reduction, while Gammaproteobacteria saw a significant

increase. Consequently, a non-excessive increase in Gram – bacteria could be used as a

maturity indicator, although the type of bacteria should be expanded and accompanied by

physicochemical and biological maturity criteria.

As previously commented, fungi are the main source of nutrition of epigeic earthworms

which means that their decrease in the presence of E. andrei is to be expected. As example,

Huang et al. (2013) observed that earthworms caused a decrease in fungal density during the

vermicomposting of fruits and vegetables. Nonetheless, after removing the earthworms the

abundance of PLFA markers for fungi increased in fresh manure, suggesting the presence of

substrates available for these organisms, accompanied by an increase in protease and

cellulase activity. The evolution of enzymes and microbial groups in the vermicompost

obtained from pig manure indicates hydrolytic activities after maturing without earthworms

for 42 days, suggesting that a greater aging time for the product may be required. Likewise,

Capítulo 4

137

fungi also increased in compost matured without earthworms, along with cellulase, protease,

and β-glucosidase. Pramanik and Chung (2011) found that the cellulolytic fungi population

increased throughout the vermicomposting of different organic waste mixes, both in the

presence of E. fetida and E. eugeniae. According to Castillo et al. (2013), the abundance of

fungi increases over the maturation period due to these microorganisms having the ability to

grow in recalcitrant compounds after a period of vermicomposting under low humidity

conditions. Also, Ryckeboer et al. (2003a) observed an increase in fungi in the maturation

process of biowaste compost.

4.3. Compost and vermicompost

According to the final parameters analyzed, the treatments with E. andrei reached

maturation and stability levels more suitable. Elvira et al. (1996) showed that the presence of

earthworms accelerates the decomposition of organic material. Consequently, the previous

step through the thermophilic phase of pig manure showed greater degradation of organic

material as a result of microbial activity and the posterior interaction of earthworms and

microbiota with the pre-composted material.

With the exception of the vermicompost obtained from fresh manure, which showed

optimal parameters close to neutrality, the pH of the remaining products achieved slightly

acidic values. Fornes et al. (2012) observed higher pH values in the vermicompost than the

combined compost-vermicompost process during the treatment of biosolids. Likewise, Villar

et al. (2016b) observed lower pH in sewage-sludge compost and vermicompost, suggesting

the accumulation of organic acids and an increase in nitrogen and phosphorus mineralization.

The high activity of acid phosphatase shows in these processes corroborates these results.

Albanell et al. (1988) suggested the formation of fulvic and humic acids during degradation

acidifies the vermicompost, whereas El Fels et al. (2014) show that pH tends to be neutral,

due to the formation of humic compounds which have buffer properties and that the pH

remains stable with an optimal value of 6-8. EC is a direct consequence of the mineralization

of organic material and, as such, the concentration of nutrients, such as nitrate. Although the

greatest degradation of organic material and carbon was in treatments with earthworms, the

compost showed high levels of EC; this could be the result of the greater level of irrigation

found in the vermicomposting system in order to maintain optimal conditions during the

process and a lower output of soluble metabolites, such as ammonium. EC values were not

greater than 3mS cm-1, making it safe to apply to the soil (Lazcano et al., 2008). The C/N

values were lower in the compost, although Fialho et al. (2010) stated that the C/N ratio is

Capítulo 4

138

not a good method for monitoring process and there is no optimal relation which

characterizes humified compost, although it is considered as a given that total C/N values

below 25-20 are necessary for compost to be considered stable or not immature. According

to Zucconi et al. (1985) germination index values greater than 80% are indicative of the

absence of phytotoxic substances for plant growth. Excepting the control, the vermicompost

and compost showed a sufficient level of maturation, the vermicompost being more

appropriate for plant growth. Atiyeh et al. (2000) showed that earthworms play an important

role in processing of cow manure, since the earthworm’s activity accelerates the

decomposition and stabilization process of manure and promotes biochemical characteristics

which are favorable to plant growth. Likewise, the compost showed a lower degree of

maturation and stability due to a respiration rate above 0.5 mg O2 g-1SV h-1 (Iannotti et al.,

1993) and an ammonium/nitrate ratio above 0.16 (Bernal et al., 1998). The inoculation of

earthworms in the maturation of pre-composted manure improved quality parameters of

compost matured under static conditions.

According to the cluster, the PLFA profiles showed that E. andrei earthworms had a

great influence on the microbial community’s structure. The final products had different

microbial communities, there being greater differences between vermicomposts, especially

that arising from fresh manure, and compost treated without earthworms. Fernández Gómez

et al. (2012) reported that vermicomposts produced from waste products of different origins

can contain similar microbial communities if the same species of earthworm is used. Sen and

Chandra (2009) showed a divergent bacterial community in compost and vermicompost

obtained from the same initial waste, despite similar changes in their physicochemical

parameters. The inoculation of E. andrei earthworms made the development of a microbial

structure different to those treatments without those detritivore organisms possible, despite

the similarity in origin, although the former composting process assimilated the microbial

communities; this established greater similarity between the products after compost

maturation without earthworms and maturation via vermicomposting. Villar et al. (2016b)

stated that the differences in microbial communities can be attributed to the microbiological

and physicochemical composition of the waste products and the interaction of the

earthworms in the microbial communities. The physicochemical composition of the initial

materials contributes to a lesser extent to separating the products, which means that the

presence or absence of earthworms determines the PLFA profile.

The structure of the microbial communities through the PLFA study is not considered

to be a quality criterion for compost or vermicompost, but it does help one to understand the

Capítulo 4

139

complex biological processes that occur during the degradation processes, which are

influenced as much by the nature of the substrate, the physicochemical changes which it

undergoes, the interaction between the microbiota and earthworms, and environmental

factors.

5. CONCLUSIONS

From a biological and final product quality standpoint, the aforementioned results

suggest that vermicomposting and compost maturation with earthworms are better

treatment options for pig manure than composting. The presence of earthworms E. andrei are

able to degrade fresh pig manure with optimal quality values, although the tracking of the

microbial community and the enzymes showed an increase of both after 42 days of

vermicompost aging, which is why we consider it necessary to optimize maturation time after

removing the earthworms. The use of earthworms to mature compost allows one to

accelerate the degradation process of the organic material, increasing the degradation of

recalcitrant compounds, in contrast with the compost matured under static conditions. The

stability of enzymatic activity, the maintenance of an abundance of fungi and bacteria, and the

study of the physicochemical parameters show that maturing compost under static

conditions produces slow and continuous degradation, which increases processing time.

Enzymatic activity is complex, which is why it is necessary to accompany them with a

physicochemical study and microbiological measurements, such as PLFAs, DGGE,

microarrays, etc.; these allow one to understand the complexity of the processes which occur

during the treatment of waste via composting and vermicomposting. Likewise, studying the

microbiology of these processes is not advisable if one does not study their activity and the

resulting physicochemical changes.

ACKNOWLEDGMENTS

This study was financially supported by the Xunta de Galicia (Regional Autonomous

Government of Galicia) (09MDS024310PR). The authors thank the research support services

of the University of Vigo (CACTI) for the carbon and nitrogen analysis.

Capítulo 4

140

REFERENCES

Aira, M., Domínguez, J., 2009. Microbial and nutrient stabilization of two animal manures after

the transit through the gut of the earthworm Eisenia fetida (Savigny, 1826). J. Hazard. Mater. 161,

1234–1238. doi:10.1016/j.jhazmat.2008.04.073

Aira, M., Monroy, F., Domínguez, J., 2007. Earthworms strongly modify microbial biomass and

activity triggering enzymatic activities during vermicomposting independently of the application rates

of pig slurry. Sci. Total Environ. 385, 252–61. doi:10.1016/j.scitotenv.2007.06.031

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.

doi:10.1007/BF00260823

Atiyeh, R.M., Domínguez, J., Subler, S., Edwards, C.A., 2000. Changes in biochemical properties of

cow manure during processing by earthworms (Eisenia andrei, Bouché) and the effects on seedling

growth. Pedobiologia 44, 709–724. doi:10.1078/S0031-4056(04)70084-0

Benitez, E., Nogales, R., Elvira, C., Masciandaro, G., Ceccanti, B., 1999. Enzyme activities as

indicators of the stabilization of sewage sludges composting with Eisenia foetida. Bioresour. Technol.

67, 297–303. doi:10.1016/S0960-8524(98)00117-5

Benítez, E., Sainz, H., Melgar, R., Nogales, R., 2002. Vermicomposting of a lignocellulosic waste

from olive oil industry: A pilot scale study. Waste Manag. Res. 20, 134–142.

doi:10.1177/0734242X0202000205

Bernal, M.P., Paredes, C., Sánchez-Monedero, M.A., Cegarra, J., 1998. Maturity and stability

parameters of composts prepared with a wide range of organic wastes. Bioresour. Technol. 63, 91–99.

doi:10.1016/S0960-8524(97)00084-9

Castillo, J.M., Romero, E., Nogales, R., 2013. Dynamics of microbial communities related to

biochemical parameters during vermicomposting and maturation of agroindustrial lignocellulose

wastes. Bioresour. Technol. 146, 345–354. doi:10.1016/j.biortech.2013.07.093

Chan, P.L., Griffiths, D., 1988. The vermicomposting of pre-treated pig manure. Biol. Wastes 24,

57–69. doi:10.1016/0269-7483(88)90027-4

Danon, M., Franke-Whittle, I.H., Insam, H., Chen, Y., Hadar, Y., 2008. Molecular analysis of

bacterial community succession during prolonged compost curing. FEMS Microbiol. Ecol. 65, 133–144.

doi:10.1111/j.1574-6941.2008.00506.x

Domínguez, J., 2004. State-of-the-Art and New Perspectives on Vermicomposting Research.

Earthworm Ecol. 401–424. doi:10.1201/9781420039719.ch20

Eivazi, F., Tabatabai, M.A., 1988. Glucosidases and galactosidases in soils. Soil Biol. Biochem. 20,

601–606. doi:10.1016/0038-0717(88)90141-1

Eivazi, F., Tabatabai, M.A., 1977. Phosphatases in soils. Soil Biol. Biochem. 9, 167–172.

doi:10.1016/0038-0717(77)90070-0

El Fels, L., Zamama, M., El Asli, A., Hafidi, M., 2014. Assessment of biotransformation of organic

matter during co-composting of sewage sludge-lignocelullosic waste by chemical, FTIR analyses, and

phytotoxicity tests. Int. Biodeterior. Biodegrad. 87, 128–137. doi:10.1016/j.ibiod.2013.09.024

Elouaqoudi, F.Z., El Fels, L., Amir, S., Merlina, G., Meddich, A., Lemee, L., Ambles, A., Hafidi, M.,

2015. Lipid signature of the microbial community structure during composting of date palm waste

alone or mixed with couch grass clippings. Int. Biodeterior. Biodegrad. 97, 75–84.

doi:10.1016/j.ibiod.2014.08.016

Capítulo 4

141

Elvira, C., Goicoechea, M., Sampedro, L., Mato, S., Nogales, R., 1996. Bioconversion of solid paper-

pulp mill sludge by earthworms. Bioresour. Technol. 57, 173–177. doi:10.1016/0960-8524(96)00065-

X

Fernández Gómez, M.J., Díaz-Raviña, M., Romero, E., Nogales, R., 2013. Recycling of

environmentally problematic plant wastes generated from greenhouse tomato crops through

vermicomposting. Int. J. Environ. Sci. Technol. 10, 697–708. doi:10.1007/s13762-013-0239-7

Fernández Gómez, M.J., Nogales, R., Insam, H., Romero, E., Goberna, M., 2012. Use of DGGE and

COMPOCHIP for investigating bacterial communities of various vermicomposts produced from

different wastes under dissimilar conditions. Sci. Total Environ. 414, 664–671.

doi:10.1016/j.scitotenv.2011.11.045

Fialho, L.L., Lopes da Silva, W.T., Milori, D.M.B.P., Simões, M.L., Martin-Neto, L., 2010.

Characterization of organic matter from composting of different residues by physicochemical and

spectroscopic methods. Bioresour. Technol. 101, 1927–1934. doi:10.1016/j.biortech.2009.10.039

Fornes, F., Mendoza-Hernández, D., García-de-la-Fuente, R., Abad, M., Belda, R.M., 2012.

Composting versus vermicomposting: A comparative study of organic matter evolution through

straight and combined processes. Bioresour. Technol. 118, 296–305.

doi:10.1016/j.biortech.2012.05.028

Garcia, C., Hernández, T., Costa, F., Ceccanti, B., Ciardi, C., 1992. Changes in ATP content , enzyme

activity and inorganic nitrogen species during composting of organic wastes. Can. J. Soil Sci. 72, 243–

253. doi:10.4141/cjss92-023

Garg, V.K., Chand, S., Chhillar, a., Yadav, a., 2005. Growth and reproduction of Eisenia foetida in

various animal wastes during vermicomposting. Appl. Ecol. Environ. Res. 3, 51–59.

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011a. Epigeic Earthworms Exert a

Bottleneck Effect on Microbial Communities through Gut Associated Processes. PLoS ONE 6, 1–9.

doi:10.1371/journal.pone.0024786

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011b. Changes in microbial community

structure and function during vermicomposting of pig slurry. Bioresour. Technol. 102, 4171–4178.

doi:10.1016/j.biortech.2010.12.057

Gómez-Brandón, M., Lazcano, C., Lores, M., Domínguez, J., 2011c. Short-term stabilization of

grape marc through earthworms. J. Hazard. Mater. 187, 291–295. doi:10.1016/j.jhazmat.2011.01.011

Gómez-Brandón, M., Lores, M., Domínguez, J., 2013. Changes in chemical and microbiological

properties of rabbit manure in a continuous-feeding vermicomposting system. Bioresour. Technol.

128, 310–6. doi:10.1016/j.biortech.2012.10.112

Gómez-Brandón, M., Lores, M., Domínguez, J., 2010. A new combination of extraction and

derivatization methods that reduces the complexity and preparation time in determining phospholipid

fatty acids in solid environmental samples. Bioresour. Technol. 101, 1348–1354.

doi:10.1016/j.biortech.2009.09.047

Gunadi, B., Edwards, C. a., 2003. The effects of multiple applications of different organic wastes

on the growth, fecundity and survival of Eisenia fetida (Savigny) (Lumbricidae). Pedobiologia 47, 321–

329. doi:10.1078/0031-4056-00196

Hjorth, M., Christensen, K.V., Christensen, M.L., Sommer, S.G., 2010. Solid–liquid separation of

animal slurry in theory and practice. A review. Agron. Sustain. Dev. 30, 153–180.

doi:10.1051/agro/2009010

Capítulo 4

142

Hothorn, T., Bretz, F., Westfall, P., 2008. Simultaneous inference in general parametric models.

Biom J 50, 346–363. doi:10.1002/bimj.200810425

Huang, K., Li, F., Wei, Y., Chen, X., Fu, X., 2013. Changes of bacterial and fungal community

compositions during vermicomposting of vegetable wastes by Eisenia foetida. Bioresour. Technol. 150,

235–241. doi:10.1016/j.biortech.2013.10.006

Iannotti, D.A., Pang, T., Toth, B.L., Elwell, D.L., Keener, H.M., Hoitink, H.A.J., 1993. A quantitative

respirometric method for monitoring compost stability. Compost Sci. Util. 1, 52–65.

doi:10.1080/1065657X.1993.10757890

Insam, H., de Bertoldi, M., 2007. Microbiology of the composting process, in: Diaz, L.F., de

Bertoldi, M., Bidlingmaier, W., Stentinford, E. (Eds.), Compost Science and Technology. Waste

Management Series. Elsevier Ltd., pp. 25–48. doi:10.1016/S1478-7482(07)80006-6

Kato, K., Miura, N., Tabuchi, H., Nioh, I., 2005. Evaluation of maturity of poultry manure compost

by phospholipid fatty acids analysis. Biol. Fertil. Soils 41, 399–410. doi:10.1007/s00374-005-0855-6

Lazcano, C., Gómez-Brandón, M., Domínguez, J., 2008. Comparison of the effectiveness of

composting and vermicomposting for the biological stabilization of cattle manure. Chemosphere 72,

1013–1019. doi:10.1016/j.chemosphere.2008.04.016

MAGRAMA, 2012. Producción y consumo sostenibles y residuos agrarios.

Monroy, F., Aira, M., Domínguez, J., 2009. Reduction of total coliform numbers during

vermicomposting is caused by short-term direct effects of earthworms on microorganisms and

depends on the dose of application of pig slurry. Sci. Total Environ. 407, 5411–5416.

doi:10.1016/j.scitotenv.2009.06.048

Nannipieri, P., Kandeler, E., Ruggiero, P., 2002. Enzyme activities and microbiological and

biochemical processes in soil, in: Burns, R.G., Dick, R.P. (Eds.), Enzymes in the Environment. New York,

pp. 1–34.

Neher, D.A., Weicht, T.R., Bates, S.T., Leff, J.W., Fierer, N., 2013. Changes in Bacterial and Fungal

Communities across Compost Recipes, Preparation Methods, and Composting Times. PLoS ONE 8,

e79512–e79512. doi:10.1371/journal.pone.0079512

Parthasarathi, K., Ranganathan, L.S., 2000. Aging effect on enzyme activities in pressmud

vermicasts of Lampito mauritii (Kinberg) and Eudrilus eugeniae (Kinberg). Biol. Fertil. Soils 30, 347–

350. doi:10.1007/s003740050014

Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D., R Core Time, 2015. nlme: linear and nonlinear mixed

effects models. R package version 3.1-119.

Pramanik, P., Chung, Y.R., 2011. Changes in fungal population of fly ash and vinasse mixture

during vermicomposting by Eudrilus eugeniae and Eisenia fetida: documentation of cellulase isozymes

in vermicompost. Waste Manag. 31, 1169–75. doi:10.1016/j.wasman.2010.12.017

R Development Core Team, 2014. R: a language and environment for statistical computing.

Ravindran, B., Wong, J.W.C., Selvam, A., Sekaran, G., 2016. Influence of microbial diversity and

plant growth hormones in compost and vermicompost from fermented tannery waste. Spec. Issue

Bioenergy Bioprod. Environ. Sustain. 217, 200–204. doi:10.1016/j.biortech.2016.03.032

Ros, M., García, C., Hernández, T., 2006. A full-scale study of treatment of pig slurry by

composting: kinetic changes in chemical and microbial properties. Waste Manag. 26, 1108–1118.

doi:10.1016/j.wasman.2005.08.008

Capítulo 4

143

Ryckeboer, J., Mergaert, J., Coosemans, J., Deprins, K., Swings, J., 2003a. Microbiological aspects of

biowaste during composting in a monitored compost bin. J. Appl. Microbiol. 94, 127–137.

doi:10.1046/j.1365-2672.2003.01800.x

Ryckeboer, J., Mergaert, J., Vaes, K., Klammer, S., De Clercq, D., Coosemans, J., Insam, H., Swings, J.,

2003b. A survey of bacteria and fungi occurring during composting and self-heating processes. Ann.

Microbiol. 53, 349–410.

Satchell, J.E., Martin, K., 1984. Phosphatase activity in earthworm faeces. Soil Biol. Biochem. 16,

191–194. doi:10.1016/0038-0717(84)90111-1

Schinner, F., von Mersi, W., 1990. Xylanase-, CM-cellulase- and invertase activity in soil: an

improved method. Soil Biol. Biochem. 22, 511–515. doi:10.1016/0038-0717(90)90187-5

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.

doi:10.1016/j.biortech.2008.07.047

Sims, G.K., Ellsworth, T.R., Mulvaney, R.L., 1995. Microscale determination of inorganic nitrogen

in water and soil extracts. Commun. Soil Sci. Plant Anal. 26, 303–316.

doi:10.1080/00103629509369298

Tiquia, S.M., Wan, J.H.C., Tam, N.F.Y., 2002. Dynamics of yard trimmings composting as

determined by dehydrogenase activity, ATP content, arginine ammonification, and nitrification

potential. Process Biochem. 37, 1057–1065. doi:10.1016/S0032-9592(01)00317-X

Villar, I., Alves, D., Garrido, J., Mato, S., 2016a. Evolution of microbial dynamics during the

maturation phase of the composting of different types of waste. Waste Manag. 54, 83–92.

doi:10.1016/j.wasman.2016.05.011

Villar, I., Alves, D., Pérez-Díaz, D., Mato, S., 2016b. Changes in microbial dynamics during

vermicomposting of fresh and composted sewage sludge. Waste Manag. 48, 409–417.

doi:10.1016/j.wasman.2015.10.011

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, 1319–26. doi:10.1016/j.biortech.2008.08.014

von Wandruszka, R., 2006. Phosphorus retention in calcareous soils and the effect of organic

matter on its mobility. Geochem. Trans. 7, 6. doi:10.1186/1467-4866-7-6

Zelles, L., 1999. Fatty acid patterns of phospholipids and lipopolysaccharides in the

characterisation of microbial communities in soil: a review. Biol. Fertil. Soils 29, 111–129.

doi:10.1007/s003740050533

Zhu, W., Yao, W., Du, W., 2016. Heavy metal variation and characterization change of dissolved

organic matter (DOM) obtained from composting or vermicomposting pig manure amended with

maize straw. Environ. Sci. Pollut. Res. 23, 12128–12139. doi:10.1007/s11356-016-6364-3

Zucconi, F., Monaco, A., Forte, M., Bertoldi, M., 1985. Phytotoxins during the stabilization of

organic matter, in: Gasser, J.K.. (Ed.), Composting of Agricultural and Other Wastes. Elsevier Applied

Science Publisher, London, pp. 73–85.

DISCUSIÓN GENERAL

Discusión

147

Los resultados aportados a lo largo de esta tesis profundizan en la dinámica microbiana

que dirige la transformación de los compuestos orgánicos que se sucede durante la fase de

maduración del proceso de compostaje y durante el vermicompostaje como proceso

alternativo o como proceso complementario de maduración del compost fresco.

La maduración del proceso de compostaje ha sido tratado en la bibliografía científica

generalmente como parte de un proceso íntegro de compostaje en estudios, por ejemplo, de

parámetros fisicoquímicos, (Eklind and Kirchmann, 2000; García et al., 1990; Garcia et al.,

1991), biológicos (Ryckeboer et al., 2003; Wong, 1985) o ambos (Benito et al., 2003;

Vuorinen and Saharinen, 1997) y, en menor medida, se han enfocado los estudios sobre

criterios microbiológicos únicamente durante la fase de maduración del compost (Danon et

al., 2008; Steger et al., 2007). Durante el compostaje en instalaciones industriales es frecuente

realizar un mayor control sobre la fase bio-oxidativa o intensiva del proceso para la cual se

emplea un sistema o tecnología específica y, por lo general, presenta una duración

determinada en función de la capacidad de tratamiento. Tras un tiempo convenido, es

habitual cambiar de posición el material y depositarlo en una zona de maduración donde el

seguimiento del proceso y las intervenciones suelen ser limitadas, por lo que, se requiere

mayor conocimiento del proceso biológico durante la estancia del compost fresco en esta

etapa de maduración, así como, conocer las posibles técnicas o aplicaciones que permitan

mejorar u optimizar el proceso y los productos obtenidos. De esta manera, los capítulos 1 y 2

del presente documento de tesis abarcan el estudio de los cambios microbiológicos y

fisicoquímicos a lo largo de la fase de maduración del compostaje. En el capítulo 1 se compara

la evolución de tres residuos de diferente origen durante la etapa de maduración (tras una

fase bio-oxidativa previa en reactor estático). Los tres residuos (municipal, industrial y

agropecuario) se han seleccionado con el fin de caracterizar de manera representativa los

cambios fisicoquímicos y microbiológicos durante la fase de maduración del proceso de

compostaje de residuos orgánicos de diferente origen. En el capítulo 2 se evalúa la

maduración dinámica con respecto a la maduración estática del residuo más energético, en

este caso, el residuo industrial, mediante la realización de volteos en función de los valores de

temperatura. En los capítulos 3 y 4 se evalúa el proceso de vermicompostaje como técnica

alternativa a la fase de maduración del compost estudiando cómo afecta a la comunidad

microbiana y a los parámetros bioquímicos la introducción de lombrices de tierra epigeas E.

andrei en residuo fresco y pre-compostado. De manera que, el capítulo 3 enfoca el estudio en

el vermicompostaje y el proceso combinado compostaje-vermicompostaje de lodo de aguas

residuales de una depuradora municipal. En el capítulo 4, se compara la maduración del

compost en condiciones estáticas frente a la maduración del compost en presencia de

Discusión

148

lombrices de tierra, incluyendo el vermicompostaje como proceso opcional al compostaje

sensu estricto. Para ello se escogió el estiércol de cerdo por ser un residuo, en general, mucho

más rico, con mayor contenido nutricional, que el lodo de depuradora municipal.

Los parámetros fisicoquímicos analizados en todos los capítulos de esta tesis,

proporcionan información convencional que permite determinar la estabilidad y madurez de

los productos finales de los diferentes procesos. Sin embargo, el estudio de los parámetros de

la dinámica microbiana permite conocer su evolución a lo largo del proceso y explica, en

función de la estabilidad biológica, las diferencias entre los productos finales y el tiempo de

proceso. Por lo tanto, la dinámica microbiana en conjunto con los parámetros fisicoquímicos

ponen de manifiesto la viabilidad de los tratamientos para producir compost y vermicompost

en condiciones adecuadas, así como, determinar las necesidades de optimización de los

procesos, lo cual permite establecer criterios para seleccionar las mejores prácticas

aplicables en una planta de tratamiento de residuos orgánicos.

Caracterización de la maduración del compost de diferentes residuos orgánicos

(capítulo 1)

De la gran variabilidad de residuos orgánicos que se generan en la actualidad se

seleccionó un residuo municipal (lodo de depuradora de aguas residuales municipales), un

residuo agropecuario (estiércol de cerdo), y un residuo industrial (lodo de depuradora de

aguas residuales de la elaboración de precocinados de pescado y cefalópodos) que se

caracterizan por su diferente origen y, por lo tanto, por las diferencias tanto fisicoquímicas

como biológicas que presentan. El estudio de la fase de maduración de estos residuos de

origen tan variado, contribuyó a aportar información útil sobre el comportamiento de sus

comunidades microbianas y su actividad de degradación de los distintos compuestos

orgánicos que componen cada residuo. El paso previo de los tres residuos por la fase más bio-

oxidativa del compostaje, en las mismas condiciones de proceso, permitió la degradación de

los compuestos más fácilmente asimilables y la generación de compuestos intermedios. Los

compost frescos son colonizados por comunidades microbianas que se caracterizan por la

composición de cada compost fresco. Así, tanto el comportamiento de las comunidades

microbianas como la evolución de las actividades enzimáticas durante la maduración fueron

condicionados por el tipo de residuo pre-compostado. A este respecto, Vargas-García et al.

(2010) mostraron que la dinámica y diversidad microbiana a lo largo del estudio de un

proceso de compostaje íntegro dependían de la naturaleza del sustrato orgánico. De la misma

manera, durante la etapa de maduración permanece la influencia ejercida por el distinto

origen que presentan los residuos; ganadero, alimentario y municipal y, por lo tanto, por las

Discusión

149

diferencias tanto fisicoquímicas como biológicas que los caracteriza. El paso a través de la

fase termófila establece variaciones en estos parámetros, sin embargo, el material pre-

compostado mantiene condiciones que diferencian, discriminan o facilitan el desarrollo de

comunidades microbianas particulares. Ishii and Takii, (2003) establecieron que los

microorganismos que proliferan a lo largo del compostaje se adaptan al medio ambiente del

sustrato y son seleccionados por factores del propio material compostado. Así, el lodo

municipal, con una diversidad microbiana inicial elevada, presentó una colonización por

parte de bacterias mesófilas y, en menor medida hongos, que dirigieron las actividades

enzimáticas del ciclo de carbono, nitrógeno y fósforo. El estiércol de cerdo se caracterizó por

una biomasa bacteriana mesofílica Gram +, elevados valores de actividad celulasa durante

toda la maduración y un aumento de las actividades fosfatasas y β-glucosidasa al final del

proceso. Las mayores diferencias en el proceso de maduración se observaron en el lodo

industrial el cual experimentó una reactivación en su estancia en maduración debido a su alto

contenido energético. El carácter lipídico y ácido, además del alto contenido en nutrientes,

que caracteriza al lodo procedente de la elaboración de alimentos precocinados, favorecieron

el desarrollo de una comunidad fúngica a lo largo de la maduración, en mayor medida que el

lodo municipal y el estiércol de cerdo. La evolución microbiana junto con los parámetros

fisicoquímicos permitieron determinar que: el estiércol de cerdo requiere mayor tiempo de

proceso u otro tipo de práctica durante la maduración; el lodo municipal alcanzó valores de

estabilidad por lo que el tiempo de proceso y el sistema de maduración se adaptaron al

residuo; y el lodo industrial requiere un cambio del proceso de maduración para tratar de

mejorar el producto y ajustar la duración del proceso. A la vista de los resultados obtenidos a

partir de la caracterización de la dinámica microbiana y la evolución de los parámetros

fisicoquímicos a lo largo de la fase de maduración, se considera necesario caracterizar los

residuos, previamente a su compostaje en planta, con la finalidad de determinar cuáles

pueden ser sus requerimientos de proceso y las mejores prácticas posibles a adaptar en las

instalaciones de tratamiento.

Influencia del diferente manejo sobre la maduración del compostaje (capítulo 2)

El residuo de la industria alimentaria experimenta un proceso de reactivación en la fase

de maduración, como consecuencia del alto contenido nutricional en compuestos energéticos

como los lípidos, por ello posibilita el estudio de la influencia de los tratamientos de manejo.

Tras la etapa intensiva en el reactor de compostaje, el material pre-compostado conservó los

suficientes nutrientes para desencadenar el crecimiento exponencial de la microbiota hasta

que su actividad propició el desarrollo de comunidades termófilas. La temperatura es uno de

Discusión

150

los principales factores que determina los cambios y sucesiones de las comunidades

microbianas y, por lo tanto, de sus actividades enzimáticas durante el compostaje. En

consecuencia, todos los grupos microbianos se vieron mayormente afectados por el régimen

de temperaturas termófilas, de modo que, las comunidades microbianas evolucionaron de

manera diferente dependiendo del sistema aplicado. Durante la maduración dinámica se

redujo la abundancia de todos los grupos microbianos mientras que esta reducción fue menor

en el tratamiento estático donde aumentó la abundancia de PLFAs biomarcadores de

bacterias Gram – y se observó un mayor ratio monoinsaturados/saturados indicativo de que

las comunidades microbianas no están sometidas a estrés y concluyendo que existe

disponibilidad suficiente de recursos. De la misma manera que en el capítulo 1, se observó la

predominancia de hongos a lo largo de la maduración como consecuencia del efecto del

residuo de origen. Sin embargo, la diferente manipulación a la que se sometió el material pre-

compostado y los cambios hacia valores de madurez provocaron una variación a

comunidades microbianas más diversas si sobre el material se realizan volteos según

criterios de proceso, en este caso, por control de la temperatura. Así, las diferentes prácticas

sobre el material pre-compostado provocaron efectos en la evolución de los grupos

microbiano. Además de los previsibles efectos que la reactivación puede presentar sobre un

proceso a escala industrial, como olores y lixiviados, conocer cómo se suceden los cambios en

la microbiota y su actividad si el material pre-compostado se mantiene en condiciones

estáticas o se voltea según criterios biológicos, permite obtener información para optimizar

el proceso de compostaje de residuos altamente energéticos. De esta manera, los estudios de

esta tesis demostraron que, si cuando se produce un cambio de posición del material pre-

compostado hacia la zona de maduración se monitoriza la temperatura y se ejercen volteos

cuando se sobrepasan valores cercanos a los 55ºC, se alarga la duración de la fase termófila.

La prolongación de las condiciones termofílicas incrementa la degradación de la materia

orgánica alcanzando criterios de maduración y estabilidad en menos tiempo que el

mantenimiento del compost en condiciones estáticas. El empleo de los volteos durante el

compostaje es una práctica esencial para el control de la temperatura y el oxígeno en

diferentes sistemas de compostaje y, en muchos casos, opcional durante la fase de

maduración en función del espacio en planta o de los recursos disponibles. Ruggieri et al.

(2008) destacaron que un residuo altamente lipídico presenta mejores resultados si su

compostaje se realiza mediante medios dinámicos que estáticos, al permitir la

homogenización y rotura de los agregados que se forman durante el compostaje. Los trabajos

de esta tesis demuestran que estos volteos son efectivos durante la fase de maduración tras

un pre-compostaje previo y controlado en un reactor con aireación forzada. Por lo que, se

Discusión

151

considera necesario prestar mayor atención a la estancia en maduración de un material pre-

compostado energético con la finalidad de mejorar el proceso de compostaje. Para una planta

de compostaje que presente una tecnología en túneles para materiales energéticos se

recomienda emplear la maduración volteada por control de temperatura u oxígeno con la

finalidad de reducir el tiempo de proceso y mejorar el producto obtenido.

Debido al auge de los productos procesados, las novedosas investigaciones realizadas

sobre este residuo exponen la viabilidad del compostaje como tratamiento de valorización.

Los resultados expuestos pueden servir como referencia para la aplicación del compostaje en

residuos de origen similar, teniendo en consideración, el aumento en la generación de este

tipo de residuos como consecuencia del incremento en la fabricación de productos

ultracongelados y procesados,

Viabilidad del empleo de lombrices E. andrei para el tratamiento de material fresco

o pre-compostado (capítulo 3 y 4)

Tanto el lodo municipal como el estiércol de cerdo presentaron una maduración en

condiciones mesófilas tras el pre-compostaje en reactor. La elevación térmica afecta a la

supervivencia de las lombrices de tierra que presentan una temperatura de desarrollo óptima

mesófila, por lo tanto, en ambos residuos pre-compostados existe la posibilidad de inocular

lombrices de tierra. De ambos residuos orgánicos existen trabajos previos donde se emplean

lombrices de tierra para su tratamiento biológico como los estudios de Aira et al. (2002)

sobre el vermicompostaje de estiércol de cerdo y Benitez et al. (1999) sobre el

vermicompostaje de lodo de depuradora, ambos con la misma especie Eisenia fetida. En

cuanto a la inoculación de lombrices de tierra en residuo pre-compostado, existen menos

trabajos de investigación, a modo de ejemplo, Chan and Griffiths (1988) realizaron pre-

compostaje de estiércol de cerdo ante la imposibilidad de vermicompostar el residuo fresco y

Rodríguez-Canché et al. (2010) pre-compostaron fangos de tanque séptico para conseguir

mantener una supervivencia máxima de las lombrices de tierra. En los estudios presentados

en esta tesis las lombrices de tierra E. andrei inoculadas en el lodo de depuradora municipal,

tanto fresco como pre-compostado, y en estiércol de cerdo, tanto fresco como pre-

compostado, presentaron valores elevados de supervivencia y, en general, parámetros

adecuados de crecimiento y desarrollo. Estos resultados constatan la viabilidad de las

lombrices de tierra epigeas para alimentarse y desarrollarse en estos residuos frescos y la

posibilidad de emplear estos organismos detritívoros para la maduración de compost fresco

como proceso complementario a la maduración del compost. Si bien, los datos mostraron

mejores condiciones para el desarrollo de estos organismos en los residuos que no fueron

Discusión

152

sometidos a compostaje. A este respecto, existen en la bibliografía investigaciones que

sustentan estos resultados como Frederickson et al. (1997), Gunadi and Edwards (2003) o

Yadav et al. (2015), quienes observaron que el pre-compostaje puede reducir los nutrientes

biodisponibles en sustratos tan variados como residuos verdes, estiércoles y residuos de

alimentos. Esta reducción del contenido nutricional en el lodo municipal junto con,

posiblemente, altos contenidos de amonio afectaron a la reproducción y crecimiento de los

organismos. En el caso del estiércol pre-compostado, los resultados mostraron que el

contenido de compuestos lábiles o volátiles pueden ser los causantes de mortalidades

iniciales, ya que una vez que estos compuestos se reducen, el contenido nutricional del

compost fresco permitió el desarrollo de las lombrices de tierra. A este respecto, puede ser

suficiente una mayor aireación del material pre-compostado a su salida del reactor, para la

reducción de los compuestos volátiles, o emplear sistemas de vermicompostaje abiertos en

lugar de sistemas cerrados, como los seleccionados para la realización de los trabajos de

investigación de esta tesis debido a que presentan un mayor control sobre las posibles

contaminaciones externas e imposibilitan la fuga de las lombrices de tierra hacia el exterior.

Determinar el sistema idóneo para cada tipo de residuo, fresco o pre-compostado, se percibe

como un paso previo importante para el diseño de procesos de vermicompostaje a mayor

escala, considerando también la introducción de material de escape y/o refugio para las

lombrices de tierra ante compuestos tóxicos o volátiles que no hayan podido eliminarse en el

acondicionamiento previo.

Efecto de E. andrei sobre la dinámica microbiana en lodo municipal y estiércol de

cerdo (capítulos 3 y 4)

Desde el punto de vista de la evolución de las características microbiológicas de los

sustratos estudiados, se observó que el lodo municipal, a pesar de presentar mayor biomasa

microbiana inicial tanto en fresco como pre-compostado, presentó limitados recursos para el

desarrollo de la microbiota mesófila Esto es consecuencia de la degradación biológica

durante el tratamiento de las aguas residuales en la estación depuradora que reduce el

contenido nutricional del lodo y, posteriormente, como consecuencia del paso del residuo por

la fase termófila del proceso de compostaje. Así, la inoculación de lombrices de tierra provocó

un descenso más rápido en las enzimas celulasa, fosfatasas y proteasa en el material pre-

compostado, mientras que en el lodo fresco las actividades enzimáticas presentaron mayor

similitud al lodo fresco sin lombrices. De modo que, se observó que las lombrices de tierra

ejercen un mayor efecto sobre las actividades degradativas en condiciones más limitantes al

aumentar la competencia por los recursos, consumir la microbiota y reducir los sustratos

Discusión

153

para las enzimas hidrolíticas. Sin embargo, el efecto sobre la biomasa microbiana fue mayor

en el lodo fresco debido a una mayor población de lombrices de tierra juveniles que

compitieron y se alimentaron de la microbiota provocando un descenso notable en las

comunidades microbianas. De esta manera, la microbiota en el lodo fresco vermicompostado

fue más eficiente en la degradación de los compuestos orgánicos al mantener actividades

enzimáticas similares al control con menor biomasa microbiana. Este efecto también se

observó en el estiércol de cerdo donde los controles presentaron una mayor carga

microbiana no siempre acompañada de una mayor actividad enzimática. Estos resultados

indican la existencia de una biomasa menos eficiente en la degradación de la materia orgánica

y mayor disponibilidad de recursos para los microorganismos en los controles, mientras que

en presencia de lombrices de tierra los microorganismos se asimilan más eficientes. Por lo

tanto, dependiendo del tipo de sustrato, puede aparecer, como consecuencia de la acción de

las lombrices de tierra epigeas, una microbiota reducida, pero metabólicamente más activa,

tal y como han señalado otros autores (Gómez-Brandón et al., 2011; Zhang et al., 2000).

Tanto en el lodo pre-compostado como en el estiércol pre-compostado, la actividad de

las lombrices redujo en mayor grado la biomasa fúngica que la bacteriana en comparación

con los controles. Sin embargo, en el lodo fresco las lombrices redujeron mayoritariamente

las bacterias Gram + y los hongos que las bacterias Gram –, mientras que en el estiércol fresco

la retirada de las lombrices de tierra provocó la recuperación de las poblaciones bacterianas

y una leve reducción de hongos en comparación con el control. Estos resultados son

congruentes con estudios previos que apuntan a un efecto selectivo por parte de las

lombrices de tierra epigeas sobre la presencia y la abundancia de determinados grupos

microbianos (Gómez-Brandón et al., 2012). Por lo tanto, las lombrices de tierra ejercen un

efecto selectivo dependiente de sustrato siendo este efecto mayor sobre la comunidad

fúngica, lo cual es debido a que los hongos son parte principal de la dieta de las lombrices de

tierra epigeas (Schonholzer et al., 1999). El estiércol de cerdo presentó altos contenidos de

moléculas complejas que no fueron sometidas a una degradación controlada previa como

ocurrió en los tanques de tratamiento biológico de la planta de aguas residuales. Así, la

biomasa microbiana, de manera más importante en el estiércol fresco, aumentó tras la

retirada de las lombrices de tierra indicando la presencia de suficientes compuestos

degradables para la microbiota. Este efecto de recuperación de las comunidades microbianas

no se observó tras la retirada de las lombrices de tierra en el lodo municipal, indicando un

mayor contenido nutricional en el estiércol. Además, las actividades enzimáticas presentaron,

en general, valores más altos que en el lodo municipal a lo largo de los distintos tratamientos

y en los productos finales, pero, en general, estas actividades hidrolíticas no se

Discusión

154

correlacionaron con los grupos microbianos. La falta de recursos en el lodo municipal

conlleva a que sea la microbiota asociada a cada sustrato la que dirija los procesos de

degradación de la materia orgánica, sin embargo, en el estiércol se sucede una interacción

lombrices-microbiota-sustrato que ejerce complejos efectos sobre las actividades

degradativas, mayor cuanto mayor contenido nutricional presente, es decir, en el proceso de

vermicompostaje con estiércol fresco.

Los vermicomposts de ambos procesos presentaron actividades biológicas y

parámetros fisicoquímicos similares en el lodo municipal y ligeramente mejores en el

vermicompost a partir de estiércol pre-compostado frente al fresco. Por lo que, tanto el

vermicompostaje del sustrato fresco como la maduración del material pre-compostado con E.

andrei son prácticas recomendables para el tratamiento del lodo de depuradora municipal y

el estiércol de cerdo.

Efecto de la inoculación de lombrices de tierra en material pre-compostado como

posible alternativa al compostaje estático de estiércol de cerdo (capítulo 4)

La maduración del compost fresco de estiércol no posibilitó su control mediante

criterios biológicos por aumento de temperatura (capítulo 1), por lo tanto, se comparó la

maduración del compost mediante condiciones estáticas frente a condiciones dinámicas

mediante la inoculación de lombrices de tierra E. andrei. Así, se observó que en el estiércol de

cerdo el compostaje previo y el posterior vermicompostaje favorecen la estabilización de los

materiales y, por lo tanto la mejora de la calidad de los productos obtenidos. El proceso de

vermicompostaje utilizando como sustrato material pre-compostado no es novedoso y ha

sido estudiado por diversos autores aunque, en la mayoría de las investigaciones, los trabajos

se basan en la determinación y seguimiento de parámetros fisicoquímicos o, por ejemplo, de

organismos patógenos. Así, Tognetti et al. (2007) observaron que el vermicompostaje de

residuo orgánico municipal, tras un compostaje previo de 50 días, conducía a un producto

final más rico en nutrientes y materia orgánica que el compostaje mientras que Mupondi et al.

(2010) aconsejan el pre-compostaje para la eliminación de patógenos previo al

vermicompostaje de estiércol. Es posible referenciar más trabajos de investigación realizados

en base a otros objetivos y sobre otro tipo de residuos, sin embargo, estudios comparativos

sobre la maduración del compost con lombrices de tierra que profundicen en la dinámica

microbiana son escasos, por ejemplo, Sen and Chandra (2009). Ambos procesos, compostaje y

vermicompostaje, son procesos biológicos que son dirigidos por los microorganismos

presentes en el sustrato y que se ven afectados tanto por factores fisicoquímicos

(temperatura, humedad, oxígeno,…) como biológicos (lombrices de tierra), por lo que el

Discusión

155

estudio de la microbiota y su actividad se considera sumamente importante para entender y

mejorar estos procesos de biodegradación.

Así, en la maduración del compost sin la adición de lombrices de tierra se observó una

relación clara entre las actividades enzimáticas y la evolución de la biomasa microbiana,

conllevando el mantenimiento estático del compost un proceso lento pero progresivo de

degradación de los compuestos orgánicos. La presencia de lombrices de tierra ejerció

complejas interacciones entre las lombrices-microbiota-sustrato afectando a las actividades

degradativas, de manera que, diversos factores provocaron cambios en las actividades

enzimáticas (actividades intracelulares o extracelulares, enzimas propias del intestino de las

lombrices, formación de complejos, etc) y no únicamente por la síntesis directa de enzimas

por parte de la microbiota presente en el sustrato. La evolución de la dinámica microbiana y

los valores fisicoquímicos del compost de estiércol de cerdo en maduración estática indicaron

que el proceso de degradación se mantuvo constante por lo que se considera necesario

establecer cambios para acelerar la maduración tratando de estimular el proceso como se

observó mediante la inoculación de lombrices de tierra epigeas. Cabe destacar que el paso del

residuo a través de la etapa termófila no únicamente presenta efectos en la abundancia de los

grupos microbianos a lo largo de la maduración sino que también en el perfil de PLFAs. De

manera que, el efecto de las lombrices de tierra epigeas sobre el material pre-compostado

varía el perfil de PLFAs pero en menor medida que sobre los residuos frescos, por lo que, el

compost y el producto combinado compost-vermicompost presentan comunidades

microbianas más similares que el vermicompost a partir de estiércol fresco. Así, Sen and

Chandra (2009) encontraron comunidades divergentes en compost y vermicompost del

mismo residuo y Lores et al. (2006) mostraron que los perfiles de las comunidades varían con

el tipo de sustrato y la especie de lombriz empleada. En este caso, se demuestra que el perfil

de la comunidad microbiana varía en mayor medida por los cambios que las lombrices

ejercen sobre el estiércol fresco mientras que el pre-compostaje asimila los perfiles de PLFAs

en los productos. Como se comentó en el apartado anterior, la retirada de lombrices de tierra

en el estiércol fresco provocó una reducción de la competencia y de depredación sobre la

microbiota, generando aumentos en las comunidades microbianas y las actividades

enzimáticas e indicando presencia de compuestos biodegradables. Por ello, es necesario

alargar el tiempo de proceso para estabilizar las actividades degradativas. A la vista de los

resultados, el estudio de la dinámica microbiana y los parámetros de calidad (ratio C/N, ratio

amonio/nitrato, índice de germinación, etc.) muestran que los vermicompost (a partir de

residuo fresco y pre-compostado) presentaron mejores condiciones que el compost y, por lo

tanto, la inoculación de lombrices de tierra en el material pre-compostado mejora el proceso

Discusión

156

de degradación y estabilización del estiércol de cerdo, siendo un proceso alternativo a la

maduración sensu estricto del compostaje.

CONCLUSIONES

Conclusiones

159

Los resultados de esta tesis muestran que las actividades enzimáticas y las

comunidades microbianas son parámetros que deben determinarse conjuntamente ya que, a

pesar de ofrecer conocimiento relevante del proceso por separado, su estudio independiente

no permite una estima adecuada del estado de degradación de la materia orgánica y, por lo

tanto, si un compost o vermicompost ha alcanzado criterios adecuados de madurez o

estabilidad. A la vista de los resultados de los distintos capítulos, no se puede establecer un

nivel base o límite de biomasa microbiana que permita distinguir un compost maduro o

inmaduro y de la misma manera, tampoco se puede establecer un valor límite o adecuado

para las actividades enzimáticas. Sin embargo, el empleo combinado de ambas estimaciones

aporta información sobre la evolución del proceso de maduración y del estado de

degradación de los compuestos orgánicos permitiendo establecer las posibles prácticas para

optimizar el proceso y el tiempo en obtener materiales más estables y maduros.

En base a los objetivos planteados inicialmente en este estudio, las siguientes

conclusiones constituyen una síntesis de los resultados más relevantes obtenidos.

1. De manera general, la comunidad microbiana que es capaz de desarrollarse en

un compost fresco en maduración depende del residuo de origen de manera que los

microorganismos mesofílicos predominantes capaces de recolonizar el compost tras las

condiciones termofílicas son similares a las comunidades microbianas predominantes del

residuo de partida. A pesar de existir un continuo cambio en las poblaciones microbianas se

mantiene la influencia del material originario y de los grupos microbianos predominantes

(capítulo 1)

2. La caracterización del residuo inicial es un proceso previo fundamental para

poder establecer una fase de maduración adaptada a cada tipo de residuo (capítulo 1).

3. El empleo de los volteos como mecanismo de control del proceso de

compostaje en residuos altamente energéticos permite un mayor grado de degradación de la

materia orgánica, mejores valores de estabilidad y madurez en menor tiempo, resultando en

una sucesión microbiana hacia comunidades mesofílicas más estables (capítulo 2).

4. El compostaje es un proceso de valorización apropiado para el lodo de

procesado de productos precocinados y ultracongelados y su estudio es un referente para

residuos con características similares.

Conclusiones

160

5. Ambas, la estructura de la comunidad microbiana y las actividades

enzimáticas, proveen importante información para el monitorio del proceso de compostaje y

el vermicompostaje. La diferente relación entre los grupos microbianos y las actividades

enzimáticas durante los distintos procesos informan del estado de degradación de la materia

orgánica y la calidad de los composts y vermicomposts, así como las necesidades de mayor o

menor tiempo de proceso o de la implantación de diferentes intervenciones sobre el material

(capítulos 1, 2, 3 y 4)

6. La inoculación de las lombrices de tierra epigeas E. andrei en todos los

residuos estudiados mostró condiciones adecuadas de supervivencia y desarrollo que ponen

de manifiesto la viabilidad de su uso para la valorización de sustratos de diferente origen

(capítulos 3 y 4).

7. Las complejas interacciones microbiota-lombrices-residuo afectan a las

actividades degradativas, por lo que el estudio de los grupos microbianos requiere del

estudio de parámetros de actividad, y viceversa, para poder concluir el devenir de los

procesos de biodegradación de los compuestos orgánicos (capítulos 3 y 4)

8. El empleo de lombrices de tierra epigeas E. andrei es recomendable para

mejorar el proceso de degradación biológica de los compuestos orgánicos en contraposición a

la maduración estática (capítulo 4).

9. Debe prestase más atención a la estancia en maduración del proceso de

compostaje y a los manejos a los que los materiales pre-compostados son sometidos,

especialmente, si se trata de residuos con alto contenido orgánico con capacidad de re-

activación. De esta manera, el proceso de compostaje debe ser diseñado y controlado en todas

sus fases, siendo posible establecer diseños ad hoc para obtener compost y vemicompost de

mejor calidad, en menor tiempo, en función del tipo de residuo y su evolución microbiana

(capítulos 1, 2, 3 y 4).

BIBLIOGRAFÍA

Bibliografía

163

A Aira, M., Domínguez, J., 2009. Microbial and nutrient stabilization of two animal

manures after the transit through the gut of the earthworm Eisenia fetida (Savigny, 1826). J. Hazard. Mater. 161, 1234–1238. doi:10.1016/j.jhazmat.2008.04.073

Aira, M., Gómez-Brandón, M., González-Porto, P., Domínguez, J., 2011. Selective reduction of the pathogenic load of cow manure in an industrial-scale continuous-feeding vermireactor. Bioresour. Technol. 102, 9633–7. doi:10.1016/j.biortech.2011.07.115

Aira, M., Monroy, F., Domínguez, J., 2007. Earthworms strongly modify microbial biomass and activity triggering enzymatic activities during vermicomposting independently of the application rates of pig slurry. Sci. Total Environ. 385, 252-261.

Aira, M., Monroy, F., Dominguez, J., 2007a. Eisenia fetida (Oligochaeta : Lumbricidae) modifies the structure and physiological capabilities of microbial communities improving carbon mineralization during vermicomposting of pig manure. Microb. Ecol. 54, 662–671. doi:10.1007/s00248-007-9223-4

Aira, M., Monroy, F., Domínguez, J., 2007b. Microbial biomass governs enzyme activity decay during aging of worm-worked substrates through vermicomposting. J. Environ. Qual. 36, 448–452. doi:10.2134/jeq2006.0262

Aira, M., Monroy, F., Domínguez, J., 2006. Eisenia fetida (Oligochaeta, Lumbricidae) activates fungal growth, triggering cellulose decomposition during vermicomposting. Microb. Ecol. 52, 738–747. doi:10.1007/s00248-006-9109-x

Aira, M., Monroy, F., Domínguez, J., Mato, S., 2002. How earthworm density affects microbial biomas and activity in pig manure. Eur. J. Soil Biol. 38, 7–10. doi:10.1016/S1164-5563(01)01116-5

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. doi:10.1007/BF00260823

Albrecht, R., Périssol, C., Ruaudel, F., Petit, J.L., Terrom, G., 2010. Functional changes in culturable microbial communities during a co-composting process: Carbon source utilization and co-metabolism. Waste Manag. 30, 764–770. doi:10.1016/j.wasman.2009.12.008

Alburquerque, J.A., Gonzálvez, J., Tortosa, G., Baddi, G.A., Cegarra, J., 2009. Evaluation of “alperujo” composting based on organic matter degradation, humification and compost quality. Biodegradation 20, 257–70. doi:10.1007/s10532-008-9218-y

Alef, K., Nannipieri, P., 1995. Methods in applied soil microbiology and biochemistry. Academic Press, London.

Allison, S.D., Vitousek, P.M., 2005. Responses of extracellular enzymes to simple and complex nutrient inputs. Soil Biol. Biochem. 37, 937–944. doi:10.1016/j.soilbio.2004.09.014

Amir, S., Abouelwafa, R., Meddich, A., Souabi, S., Winterton, P., Merlina, G., Revel, J.-C., Pinelli, E., Hafidi, M., 2010. PLFAs of the microbial communities in composting mixtures of agro-industry sludge with different proportions of household waste. Int. Biodeterior. Biodegrad. 64, 614–621. doi:10.1016/j.ibiod.2010.01.012

Bibliografía

164

APHA, AWWA, WEF. 2012. Standard Methods for the Examination of Water and Wastewater. Stand Methods. doi:10.2105/AJPH.51.6.940-a

Atiyeh, R.M., Domínguez, J., Subler, S., Edwards, C.A., 2000. Changes in biochemical properties of cow manure during processing by earthworms (Eisenia andrei, Bouché) and the effects on seedling growth. Pedobiologia 44, 709–724. doi:10.1078/S0031-4056(04)70084-0

B Bastida F, Kandeler E, Moreno JL, Ros M, García C, Hernández T., 2008. Application of

fresh and composted organic wastes modifies structure, size and activity of soil microbial community under semiarid climate. Appl Soil Ecol. 40: 318–329. doi:10.1016/j.apsoil.2008.05.007

Barrington, S., Choinière, D., Trigui, M., Knight, W., 2002. Effect of carbon source on compost nitrogen and carbon losses. Bioresour. Technol. 83, 189–194. doi:10.1016/S0960-8524(01)00229-2

Beffa, T., Blanc, M., Marilley, L., Fischer, J.L., Lyon, P.-F., Aragno, M., 1996. Taxonomic and metabolic microbial diversity during composting, in: de Bertoldi, M., Sequi, P., Lemmes, B., Papi, T. (Eds.), The Science of Composting. Springer Netherlands, Dordrecht, pp. 149–161.

Benckiser, G. (Ed.), 1997. Fauna in soil ecosystems: recycling processes, nutrient fluxes, and agricultural production. CRC Press, New York.

Benitez, E., Elvira, C., Gomez, M., Gallardo-Lara, F., Nogales, R., 1996. Leachates from a vermicomposting process, in: Fertilizers and Environment. Springer Netherlands, Dordrecht, pp. 323–326. doi:10.1007/978-94-009-1586-2_54

Benitez, E., Nogales, R., Elvira, C., Masciandaro, G., Ceccanti, B., 1999a. Enzyme and Earthworm Activities during Vermicomposting of Carbaryl-Treated Sewage Sludge. J. Environ. Qual. 28, 1099–1104. doi:10.2134/jeq1999.00472425002800040006x

Benitez, E., Nogales, R., Elvira, C., Masciandaro, G., Ceccanti, B., 1999b. Enzyme activities as indicators of the stabilization of sewage sludges composting with Eisenia foetida. Bioresour. Technol. 67, 297–303. doi:10.1016/S0960-8524(98)00117-5

Benítez, E., Sainz, H., Melgar, R., Nogales, R., 2002. Vermicomposting of a lignocellulosic waste from olive oil industry: A pilot scale study. Waste Manag. Res. 20, 134–142. doi:10.1177/0734242X0202000205

Benitez, E., Sainz, H., Nogales, R., 2005. Hydrolytic enzyme activities of extracted humic substances during the vermicomposting of a lignocellulosic olive waste. Bioresour. Technol. 96, 785–790. doi:10.1016/j.biortech.2004.08.010

Benito, M., Masaguer, A., Moliner, A., Arrigo, N., Palma, R.M., 2003. Chemical and microbiological parameters for the characterisation of the stability and maturity of pruning waste compost. Biol. Fertil. Soils 37, 184–189. doi:10.1007/s00374-003-0584-7

Bernal, M.P., Alburquerque, J.A., Moral, R., 2009. Composting of animal manures and chemical criteria for compost maturity assessment. A review. Bioresour. Technol. 100, 5444–5453. doi:10.1016/j.biortech.2008.11.027

Bibliografía

165

Bernal MP, Paredes C, Sánchez-Monedero MA, Cegarra J., 1998. Maturity and stability parameters of composts prepared with a wide range of organic wastes. Bioresour Technol. 63: 91–99. doi:10.1016/S0960-8524(97)00084-9

Bhat, S.A., Singh, J., Vig, A.P., 2013. Vermiremediation of dyeing sludge from textile mill with the help of exotic earthworm Eisenia fetida Savigny. Environ. Sci. Pollut. Res. 20, 5975–5982. doi:10.1007/s11356-013-1612-2

Blouin, M., Hodson, M.E., Delgado, E.A., Baker, G., Brussaard, L., Butt, K.R., Dai, J., Dendooven, L., Peres, G., Tondoh, J.E., Cluzeau, D., Brun, J.J., 2013. A review of earthworm impact on soil function and ecosystem services. Eur. J. Soil Sci. 64, 161–182. doi:10.1111/ejss.12025

Boer, W. de, Folman, L.B., Summerbell, R.C., Boddy, L., 2005. Living in a fungal world: impact of fungi on soil bacterial niche development. FEMS Microbiol. Rev. 29, 795–811. doi:10.1016/j.femsre.2004.11.005

Boletín Oficial del Estado, 2013. Real Decreto 506/2013, de 28 de junio, sobre productos fertilizantes., núm. 164.

Bossio DA, Scow KM. 1998. Impacts of carbon and flooding on soil microbial communities: phospholipid fatty acid profiles and substrate utilization patterns. Microb Ecol. 35: 265–278. doi:10.1007/s002489900082

Bouché, M., 1972. Lombriciens de France: écologie et systématique, Annales de Zoologie Ecologie Animale. Paris.

Bouché, M.B., 1977. Strategies lombriciennes. Ecol. Bull. 25, 122–132. doi:10.2307/20112572

Boulter-Bitzer, J.I., Trevors, J.T., Boland, G.J., 2006. A polyphasic approach for assessing maturity and stability in compost intended for suppression of plant pathogens. Appl. Soil Ecol. 34, 65–81. doi:10.1016/j.apsoil.2005.12.007

Brito LM, Coutinho J, Smith SR., 2008. Methods to improve the composting process of the solid fraction of dairy cattle slurry. Bioresour Technol. 99: 8955–60. doi:10.1016/j.biortech.2008.05.005

Brown, G.G., 1995. How do earthworms affect microfloral and faunal community diversity? Plant Soil 170, 209–231. doi:10.1007/BF02183068

Brown, G.G., Doube, B.M., 2004. Functional interactions between earthworms, microorganisms, organic matter, and plants, in: Earthworm Ecology. St. Lucie Press, Boca Raton, Florida, pp. 213–239.

Burns, R.G., 1982. Enzyme activity in soil: Location and a possible role in microbial ecology. Soil Biol. Biochem. 14, 423–427. doi:10.1016/0038-0717(82)90099-2

Bustamante, M.A., Restrepo, A.P., Alburquerque, J.A., Pérez-Murcia, M.D., Paredes, C., Moral, R., Bernal, M.P., 2013. Recycling of anaerobic digestates by composting: effect of the bulking agent used. J. Clean. Prod. 47, 61–69. doi:10.1016/j.jclepro.2012.07.018

Bibliografía

166

C Cabanas-Vargas, D., Stentiford, E., 2006. Oxygen and CO2 profiles and methane

formation during the maturation phase of composting. Compost Sci. Util. 14, 86–89.

Cabrera, M.L., Beare, M.H., 1993. Alkaline persulfate oxidation for determining total nitrogen in microbial biomass extracts. Soil Sci. Soc. Am. J. 57, 1007–1012. doi:10.2136/sssaj1993.03615995005700040021x

Cahyani VR, Matsuya K, Asakawa S, Kimura M., 2003. Succession and phylogenetic composition of bacterial communities responsible for the composting process of rice straw estimated by PCR-DGGE analysis. Soil Sci Plant Nutr. 49: 619–630. doi:10.1080/00380768.2003.10410052

Campitelli, P., Ceppi, S., 2008. Effects of composting technologies on the chemicals and physicochemical properties of humic acids. Geoderma 144, 325–333.

Castaldi, P., Garau, G., Melis, P., 2008. Maturity assessment of compost from municipal solid waste through the study of enzyme activities and water-soluble fractions. Waste Manag. 28, 534–540. doi:10.1016/j.wasman.2007.02.002

Castillo, J.M., Romero, E., Nogales, R., 2013. Dynamics of microbial communities related to biochemical parameters during vermicomposting and maturation of agroindustrial lignocellulose wastes. Bioresour. Technol. 146, 345–354. doi:10.1016/j.biortech.2013.07.093

Cayuela, M.L., Mondini, C., Sánchez-Monedero, M.A., Roig, A., 2008. Chemical properties and hydrolytic enzyme activities for the characterisation of two-phase olive mill wastes composting. Bioresour. Technol. 99, 4255–4262. doi:10.1016/j.biortech.2007.08.057

Chang, J.I., Chen, Y.J., 2010. Effects of bulking agents on food waste composting. Bioresour. Technol. 101, 5917–5924. doi:10.1016/j.biortech.2010.02.042

Chan, P.L., Griffiths, D., 1988. The vermicomposting of pre-treated pig manure. Biol. Wastes 24, 57–69. doi:10.1016/0269-7483(88)90027-4

Chen, Y., Zhang, Q., Zhang, Y., Chen, J., Zhang, D., Tong, J., 2015a. Changes in fibrolytic enzyme activity during vermicomposting of maize stover by an anecic earthworm Amynthas hupeiensis. Polym. Degrad. Stab. 120, 169–177. doi:10.1016/j.polymdegradstab.2015.06.018

Chen, Y., Zhang, Y., Zhang, Q., Xu, L., Li, R., Luo, X., Zhang, X., Tong, J., 2015b. Earthworms modify microbial community structure and accelerate maize stover decomposition during vermicomposting. Environ. Sci. Pollut. Res. 22, 17161–17170. doi:10.1007/s11356-015-4955-z

Choi MH, Park YH., 1998. The influence of yeast on thermophilic composting of food waste. Lett Appl Microbiol. 26: 175–178. doi:10.1046/j.1472-765X.1998.00307.x

Ciavatta, C., Govi, M., Pasotti, L., Sequi, P., 1993. Changes in organic matter during stabilization of compost from municipal solid wastes. Bioresour. Technol. 43, 141–145. doi:10.1016/0960-8524(93)90173-9

Coleman, D., Crossley, D., Hendrix, P., 2004. Fundamentals of soil ecology. Academic Press, San Diego.

Bibliografía

167

Cook KL, Ritchey EL, Loughrin JH, Haley M, Sistani KR, Bolster CH., 2015. Effect of turning frequency and season on composting materials from swine high-rise facilities. Waste Manag. doi:10.1016/j.wasman.2015.02.019

Coyne, M.S., 1999. Soil microbiology: an exploratory approach. Delmar New York, NY, USA.

Curry, J.P., Schmidt, O., 2007. The feeding ecology of earthworms - A review. Pedobiologia 50, 463–477. doi:10.1016/j.pedobi.2006.09.001

Curtin, D., Campbell, C.A., Jalil, A., 1998. Effects of acidity on mineralization: pH-dependence of organic matter mineralization in weakly acidic soils. Soil Biol. Biochem. 30, 57–64. doi:10.1016/S0038-0717(97)00094-1

D Danon, M., Franke-Whittle, I.H., Insam, H., Chen, Y., Hadar, Y., 2008. Molecular analysis

of bacterial community succession during prolonged compost curing. FEMS Microbiol. Ecol. 65, 133–144. doi:10.1111/j.1574-6941.2008.00506.x

de Bertoldi, M., Sequi, P., Lemmes, B., Papi, T. (Eds.), 1996. The Science of Composting. Springer Netherlands, Dordrecht. doi:10.1007/978-94-009-1569-5

de Bertoldi, M., Vallini, G., Pera, A., 1983. The Biology of Composting: a Review. Waste Manag. Res. 157–176.

de Guardia, A., Petiot, C., Rogeau, D., Druilhe, C., 2008. Influence of aeration rate on nitrogen during blackwater composting. Waste Manag. 248, 575–587. doi:10.1016/j.wasman.2007.02.007

Diaz, L.F., Savage, G.M., 2007. Factors that affect the process, in: Diaz, L.F., De Bertoldi, M., Bidlingmaier, W., Stentinford, E. (Eds.), Compost Science and Technology, Waste Management Series. Amsterdam, pp. 49–65.

Diaz, L.F., Savage, G.M., Eggerth, L.L., Chiumenti, A., 2007. Systems used in composting, in: Compost Science and Technology. pp. 67–87.

Diaz, L.F., Savage, G.M., Golueke, C.G., 2002. Composting of municipal solid wastes, in: Tchobanoglous, G., Kreith, F. (Eds.), Handbook of Solid Waste Management. McGraw-Hill Inc., New York, p. 12.1-12.70.

Domínguez, J., 2011. The Microbiology of Vermicomposting, in: Edwards, C.A., Arancon, N.Q., Sherman, R. (Eds.), Vermiculture Technology: Earthworms, Organic Wastes and Environmental Management. CRC Press, Boca Raton, Florida. pp. 53–66.

Domínguez, J., 2004. State-of-the-Art and New Perspectives on Vermicomposting Research. Earthworm Ecol. 401–424. doi:doi:10.1201/9781420039719.ch20

Domínguez, J., Aira, M., Gómez-Brandón, M., 2010. Vermicomposting: earthworms enhance the work of microbes, in: Insam, H., Franke-Whittle, I.H., Goberna, M. (Eds.), Microbes at Work: From Wastes to Resources. Springer, Heidelberg, pp. 93–114.

Bibliografía

168

Domínguez, J., Edwards, C. A., 2011. Biology and Ecology of Earthworm Species Used for Vermicomposting, in: Edwards, C.A., Arancon, N.Q., Sherman, R. (Eds.), Vermiculture Technology: Earthworms, Organic Wastes and Environmental Management. CRC Press, Boca Raton, Florida, pp. 27–40.

Domínguez, J., Edwards, C.A., 1997. Effects of stocking rate and moisture content on the growth and maturation of Eisenia andrei (Oligochaeta) in pig manure. Soil Biol. Biochem. 29, 743–746.

Domínguez, J., Edwards, C.A., Webster, M., 2000. Vermicomposting of sewage sludge: Effect of bulking materials on the growth and reproduction of the earthworm Eisenia andrei. Pedobiologia 44, 24–32. doi:10.1078/S0031-4056(04)70025-6

Domínguez, J., Gómez-Brandón, M., 2013. The influence of earthworms on nutrient dynamics during the process of vermicomposting. Waste Manage. Res. 31, 859-868.

Domínguez, J., Pérez-Díaz, D., 2011. Desarrollo y nuevas perspectivas del Vermicompostaje, in: Gestión de Residuos Orgánicos de Uso Agrícola. Servizo de Publicacións e Intercambio Científico, Universidade de Santiago de Compostela, pp. 201–214.

Domínguez, J., Velando, A., Ferreiro, A., 2005. Are Eisenia fetida (Savigny, 1826) and Eisenia andrei. Pedobiologia 49, 81–87. doi:10.1016/j.pedobi.2004.08.005

E Eastman, B.R., Kane, P.N., Edwards, C.A., Trytek, L., Gunadi, B., Stermer, A.L., Mobley, J.R.,

2001. The Effectiveness of Vermiculture in Human Pathogen Reduction for USEPA Biosolids Stabilization. Compost Sci. Util. 9, 38–49. doi:10.1080/1065657X.2001.10702015

Edwards, C.A., 2010. Low-technology vermicomposting systems, in: Edwards, C.A., Arancon, N.Q., Sherman, R. (Eds.), Vermiculture Technology: Earthworms, Organic Wastes and Environmental Management. CRC Press, Boca Raton, Florida, pp. 79–90.

Edwards, C.A., 2004. The importance of earthworms as key representatives of the soil fauna, in: Earthworm Ecology. CRC Press, Boca Raton, Florida, pp. 3–12.

Edwards, C.A., 1988. Breakdown of animal, vegetable and industrial organic wastes by earthworms, in: Edwards, Clive A., Neuhauser, E.F. (Eds.), Earthworms in Waste and Environmental Management. SPB Academic Publishing, the Hague, pp. 21–31.

Edwards, C.A., Arancon, N.Q., 2004. The use of earthworms in the breakdown of organic wastes to produce vermicomposts and animal feed protein, in: Earthworm Ecology. St. Lucie Press, Boca Raton, Florida, pp. 345–379.

Edwards, C.A., Bohlen, P.J., 1996. Biology and ecology of earthworms. Chapman and Hall, London.

Edwards, C.A., Domínguez, J., Neuhauser, E.F., 1998. Growth and reproduction of Perionyx excavatus (Perr.) (Megascolecidae) as factors in organic waste management. Biol. Fertil. Soils 27, 155–161. doi:10.1007/s003740050414

Bibliografía

169

Edwards, C.A., Fletcher, K.E., 1988. Interactions between earthworms and microorganisms in organic-matter breakdown. Agric. Ecosyst. Environ. 24, 235–247. doi:10.1016/0167-8809(88)90069-2

Eiland, F., Klamer, M., Lind, A.-M., Leth, M., Bååth, E., 2001. Influence of initial C/N ratio on chemical and microbial composition during long term composting of straw. Microb. Ecol. 41, 272–280. doi:10.1007/s002480000071

Eivazi, F., Tabatabai, M.A., 1988. Glucosidases and galactosidases in soils. Soil Biol. Biochem. 20, 601–606. doi:10.1016/0038-0717(88)90141-1

Eivazi, F., Tabatabai, M.A., 1977. Phosphatases in soils. Soil Biol. Biochem. 9, 167–172. doi:10.1016/0038-0717(77)90070-0

Eklind, Y., Kirchmann, H., 2000. Composting and storage of organic household waste with different litter amendments. II: nitrogen turnover and losses. Bioresour. Technol. 74, 125–133. doi:10.1016/S0960-8524(00)00005-5

El Fels, L., Zamama, M., El Asli, A., Hafidi, M., 2014. Assessment of biotransformation of organic matter during co-composting of sewage sludge-lignocelullosic waste by chemical, FTIR analyses, and phytotoxicity tests. Int. Biodeterior. Biodegrad. 87, 128–137. doi:10.1016/j.ibiod.2013.09.024

Elouaqoudi, F.Z., El Fels, L., Amir, S., Merlina, G., Meddich, A., Lemee, L., Ambles, A., Hafidi, M., 2015. Lipid signature of the microbial community structure during composting of date palm waste alone or mixed with couch grass clippings. Int. Biodeterior. Biodegradation 97, 75–84. doi:10.1016/j.ibiod.2014.08.016

Elvira, C., Goicoechea, M., Sampedro, L., Mato, S., Nogales, R., 1996. Bioconversion of solid paper-pulp mill sludge by earthworms. Bioresour. Technol. 57, 173–177. doi:10.1016/0960-8524(96)00065-X

Elvira, C., Mato, S., Nogales, R., 1995. Changes in heavy metal extractability and organic matter fractions after vermicomposting of sludges from a paper mill industry and wastewater treatment plant. Fresenius Environ. Bull. 4, 503–507.

Elvira, C., Sampedro, L., Benítez, E., Nogales, R., 1998. Vermicomposting of sludges from paper mill and dairy industries with Eisena andrei: A pilot-scale study. Bioresour. Technol. 63, 205–211. doi:10.1016/S0960-8524(97)00145-4

Elvira, C., Sampedro, L., Dominguez, J., Mato, S., 1997. Vermicomposting of Wastewater Sludge from Paper-Pulp Industry with Nitrogen Rich Materials. Soil Biol. Biochem. 29, 759–762. doi:10.1016/S0038-0717(96)00202-7

Epstein, E., 2011. Industrial composting. Environ. Eng. Facil. Manag. Taylor Francis Group LLC.

Epstein, E., 1996. The science of composting. CRC press.

European Commission, 2001. Working document: biological treatment of biowaste, 2nd draft. Dir. Gen. Environ. 22–22.

Bibliografía

170

Eurostat, 2017. Waste statistics - Statistics Explained [WWW Document]. URL http://ec.europa.eu/eurostat/statistics-explained/index.php/Waste_statistics (accessed 5.10.17).

F Federici, E., Pepi, M., Esposito, A., Scargetta, S., Fidati, L., Gasperini, S., Cenci, G., Altieri,

R., 2011. Two-phase olive mill waste composting: Community dynamics and functional role of the resident microbiota. Bioresour. Technol. 102, 10965–10972. doi:10.1016/j.biortech.2011.09.062

Fernandes F, Viel M, Sayag D, André L. 1988. Microbial breakdown of fats through in-vessel co-composting of agricultural and urban wastes. Biol Wastes. 26: 33–48. doi:http://dx.doi.org/10.1016/0269-7483(88)90147-4

Fernández Gómez, M.J., Díaz-Raviña, M., Romero, E., Nogales, R., 2013. Recycling of environmentally problematic plant wastes generated from greenhouse tomato crops through vermicomposting. Int. J. Environ. Sci. Technol. 10, 697–708. doi:10.1007/s13762-013-0239-7

Fernández Gómez, M.J., Nogales, R., Insam, H., Romero, E., Goberna, M., 2012. Use of DGGE and COMPOCHIP for investigating bacterial communities of various vermicomposts produced from different wastes under dissimilar conditions. Sci. Total Environ. 414, 664–671. doi:10.1016/j.scitotenv.2011.11.045

Fernández Gómez, M.J., Nogales, R., Insam, H., Romero, E., Goberna, M., 2010a. Continuous-feeding vermicomposting as a recycling management method to revalue tomato-fruit wastes from greenhouse crops. Waste Manag. 30, 2461–2468. doi:10.1016/j.wasman.2010.07.005

Fernández Gómez, M.J., Romero, E., Nogales, R., 2011. Impact of imidacloprid residues on the development of Eisenia fetida during vermicomposting of greenhouse plant waste. J. Hazard. Mater. 192, 1886–1889. doi:10.1016/j.jhazmat.2011.06.077

Fernández Gómez, M.J., Romero, E., Nogales, R., 2010b. Feasibility of vermicomposting for vegetable greenhouse waste recycling. Bioresour. Technol. 101, 9654–9660. doi:10.1016/j.biortech.2010.07.109

Fialho, L.L., Lopes da Silva, W.T., Milori, D.M.B.P., Simões, M.L., Martin-Neto, L., 2010. Characterization of organic matter from composting of different residues by physicochemical and spectroscopic methods. Bioresour. Technol. 101, 1927–1934. doi:10.1016/j.biortech.2009.10.039

Fornes, F., Mendoza-Hernández, D., García-de-la-Fuente, R., Abad, M., Belda, R.M., 2012. Composting versus vermicomposting: A comparative study of organic matter evolution through straight and combined processes. Bioresour. Technol. 118, 296–305. doi:10.1016/j.biortech.2012.05.028

Frederickson, J., Butt, K.R., Morris, R.M., Daniel, C., 1997. Combining vermiculture with traditional green waste composting systems. Soil Biol. Biochem. 29, 725–730. doi:10.1016/S0038-0717(96)00025-9

Bibliografía

171

Frostegård, Å., Bååth, E., 1996. The use of phospholipid fatty acid analysis to estimate bacterial and fungal biomass in soil. Biol. Fertil. Soils 22, 59-65.

Fu, X., Huang, K., Chen, X., Li, F., Cui, G., 2015. Feasibility of vermistabilization for fresh pelletized dewatered sludge with earthworms Bimastus parvus. Bioresour. Technol. 175, 646-650.

G Gajalakshmi, S., Ramasamy, E.V., Abbasi, S.A., 2002. Vermicomposting of paper waste

with the anecic earthworm Lampito mauritii Kinberg. Indian J. Chem. Technol. 9, 306–311.

Garcia, C., Hernández, T., Costa, F., 1991. Changes in carbon fractions during composting and maturation of organic wastes. Environ. Manage. 15, 433–439. doi:10.1007/BF02393889

García, C., Hernández, T., Costa, F., 1990. The influence of composting and maturation processes on the heavy-metal extractability from some organic wastes. Biol. Wastes 31, 291–301. doi:10.1016/0269-7483(90)90086-8

Garcia, C., Hernández, T., Costa, F., Ceccanti, B., Ciardi, C., 1992. Changes in ATP content , enzyme activity and inorganic nitrogen species during composting of organic wastes. Can. J. Soil Sci. 72, 243–253. doi:10.4141/cjss92-023

García, C., Hernández, T., Costa, C., Ceccanti, B., Masciandaro, G., Ciardi, C., 1993. A study of biochemical parameters of composted and fresh municipal wastes. Bioresour. Technol. 44, 17–23. doi:10.1016/0960-8524(93)90202-M

García-Morales JL, Álvarez CJ, Paredes C, López E, Fernández FJ, Bustamante MA, et al. 2015. Residuos agroalimentarios I.3. Moreno J, Moral R, García-Morales JL, Pascual JA, Bernal MP (Eds). Madrid: Mundi-Prensa.

Garg, V.K., Chand, S., Chhillar, A., Yadav, A., 2005. Growth and reproduction of Eisenia foetida in various animal wastes during vermicomposting. Appl. Ecol. Environ. Res. 3, 51–59.

Garg, P., Gupta, A., Satya, S., 2006. Vermicomposting of different types of waste using Eisenia foetida: A comparative study. Bioresour. Technol. 97, 391–395. doi:10.1016/j.biortech.2005.03.009

Garg, V.K., Kaushik, P., 2005. Vermistabilization of textile mill sludge spiked with poultry droppings by an epigeic earthworm Eisenia foetida. Bioresour. Technol. 96, 1063–1071. doi:10.1016/j.biortech.2004.09.003

Gea, T., Barrena, R., Artola, A., Sánchez, A., 2007. Optimal bulking agent particle size and usage for heat retention and disinfection in domestic wastewater sludge composting. Waste Manag. 27, 1108–1116. doi:10.1016/j.wasman.2006.07.005

Gea T, Ferrer P, Alvaro G, Valero F, Artola A, Sánchez A., 2007. Co-composting of sewage sludge: fats mixtures and characteristics of the lipases involved. Biochem Eng J. 33: 275–283. doi:10.1016/j.bej.2006.11.007

Bibliografía

172

Geisseler, D., Horwath, W.R., 2008. Regulation of extracellular protease activity in soil in response to different sources and concentrations of nitrogen and carbon. Soil Biol. Biochem. 40, 3040-3048.

Getahun T, Nigusie A, Entele T, Gerven T Van, Bruggen B Van der, 2012. Effect of turning frequencies on composting biodegradable municipal solid waste quality. Resour Conserv Recycl. 65: 79–84. doi:10.1016/j.resconrec.2012.05.007

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011a. Changes in microbial community structure and function during vermicomposting of pig slurry. Bioresour. Technol. 102, 4171–4178. doi:10.1016/j.biortech.2010.12.057

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011b. Epigeic Earthworms Exert a Bottleneck Effect on Microbial Communities through Gut Associated Processes. PLoS ONE 6, 1–9. doi:10.1371/journal.pone.0024786

Gómez-Brandón, M., Lazcano, C., Lores, M., Domínguez, J., 2011. Short-term stabilization of grape marc through earthworms. J. Hazard. Mater. 187, 291–295. doi:10.1016/j.jhazmat.2011.01.011

Gómez-Brandón, M., Aira, M., Lores, M., Domínguez, J., 2011. Epigeic Earthworms Exert a Bottleneck Effect on Microbial Communities through Gut Associated Processes. PLoS ONE 6, 1–9. doi:10.1371/journal.pone.0024786

Gómez-Brandón, M., Lores, M., Domínguez, J., 2010. A new combination of extraction and derivatization methods that reduces the complexity and preparation time in determining phospholipid fatty acids in solid environmental samples. Bioresour. Technol. 101, 1348–1354. doi:10.1016/j.biortech.2009.09.047

Gómez-Brandón, M., Lores, M., Domínguez, J., 2013. Changes in chemical and microbiological properties of rabbit manure in a continuous-feeding vermicomposting system. Bioresour. Technol. 128, 310–6. doi:10.1016/j.biortech.2012.10.112

Gómez-Brandón, M., Lores, M., Domínguez, J., 2012. Species-specific effects of epigeic earthworms on microbial community structure during first stages of decomposition of organic matter. PLoS ONE 7, 1–9. doi:10.1371/journal.pone.0031895

Green SJ, Michel FC, Hadar Y, Minz D. Similarity of bacterial communities in sawdust- and straw-amended cow manure composts., 2004. FEMS Microbiol Lett. 233: 115–123. doi:10.1016/j.femsle.2004.01.049

Gunadi, B., Blount, C., Edwards, C.A., 2002. The growth and fecundity of Eisenia fetida (Savigny) in cattle solids pre-composted for different periods. Pedobiologia 23, 15–23. doi:10.1078/0031-4056-00109

Gunadi, B., Edwards, C. A., 2003. The effects of multiple applications of different organic wastes on the growth, fecundity and survival of Eisenia fetida (Savigny) (Lumbricidae). Pedobiologia 47, 321–329. doi:10.1078/0031-4056-00196

Bibliografía

173

H Haimi, J., Huhta, V., 1986. Capacity of various organic residues to support adequate

earthworm biomass for vermicomposting. Biol. Fertil. Soils 2, 23–27. doi:10.1007/BF00638957

Hait, S., Tare, V., 2011. Vermistabilization of primary sewage sludge. Bioresour. Technol. 102, 2812-2820.

Hamoda, M., Qdais, H.A., Newham, J., 1998. Evaluation of municipal solid waste composting kinetics. Resour. Conserv. Recycl. 23, 209–223.

Hartenstein, R., Hartenstein, F., 1981. Physicochemical Changes Effected in Activated Sludge by the Earthworm Eisenia foetida1. J. Environ. Qual. 10, 377–377. doi:10.2134/jeq1981.00472425001000030027x

Hassen, A., Belguith, K., Jedidi, N., Cherif, A., Cherif, M., Boudabous, A., 2001. Microbial characterization during composting of municipal solid waste. Bioresour. Technol. 80, 217–225. doi:10.1016/S0960-8524(01)00065-7

Haug, R.T., 1993. The Practical Handbook of Compost Engineering. Lewis Publishers, Boca Raton.

Hellmann, B., Zelles, L., Palojärvi, A., Bai, Q., 1997. Emission of climate-relevant trace gases and succession of microbial communities during open-windrow composting. Appl. Environ. Microbiol. 63, 1011–1018.

Hénault-Ethier, L., Martin, V.J.J., Gélinas, Y., 2016. Persistence of Escherichia coli in batch and continuous vermicomposting systems. Waste Manag. 56, 88–99. doi:10.1016/j.wasman.2016.07.033

Hjorth, M., Christensen, K.V., Christensen, M.L., Sommer, S.G., 2010. Solid–liquid separation of animal slurry in theory and practice. A review. Agron. Sustain. Dev. 30, 153–180. doi:10.1051/agro/2009010

Hothorn, T., Bretz, F., Westfall, P., 2008. Simultaneous inference in general parametric models. Biom. J. 50, 346–363. doi:10.1002/bimj.200810425

Huang, K., Li, F., Wei, Y., Chen, X., Fu, X., 2013. Changes of bacterial and fungal community compositions during vermicomposting of vegetable wastes by Eisenia foetida. Bioresour. Technol. 150, 235–241. doi:10.1016/j.biortech.2013.10.006

Huang, K., Li, F., Wei, Y., Fu, X., Chen, X., 2014. Effects of earthworms on physicochemical properties and microbial profiles during vermicomposting of fresh fruit and vegetable wastes. Bioresour. Technol. 170, 45–52. doi:10.1016/j.biortech.2014.07.058

Huet, J., Druilhe, C., Trémier, A., Benoist, J.C., Debenest, G., 2012. The impact of compaction, moisture content, particle size and type of bulking agent on initial physical properties of sludge-bulking agent mixtures before composting. Bioresour. Technol. 114, 428–436. doi:10.1016/j.biortech.2012.03.031

Bibliografía

174

Hultman, J., Vasara, T., Partanen, P., Kurola, J., Kontro, M.H., Paulin, L., Auvinen, P., Romantschuk, M., 2010. Determination of fungal succession during municipal solid waste composting using a cloning-based analysis. J. Appl. Microbiol. 108, 472–487. doi:10.1111/j.1365-2672.2009.04439.x

I Iannotti, D.A., Pang, T., Toth, B.L., Elwell, D.L., Keener, H.M., Hoitink, H.A.J., 1993. A

quantitative respirometric method for monitoring compost stability. Compost Sci. Util. 1, 52–65. doi:10.1080/1065657X.1993.10757890

Iannotti, E.L., Porter, J.H., Fischer, J.R., Sievers, D.M., 1979. Changes in swine manure during anaerobic digestion. Dev. Ind. Microbiol. 20, 519–529.

Iglesias Jiménez, E., Pérez García, V., 1991. Composting of domestic refuse and sewage sludge. I. Evolution of temperature, pH, C/N ratio and cation-exchange capacity. Resour. Conserv. Recycl. 6, 45–60. doi:10.1016/0921-3449(91)90005-9

Insam, H., de Bertoldi, M., 2007. Microbiology of the composting process, in: Diaz, L.F., de Bertoldi, M., Bidlingmaier, W., Stentinford, E. (Eds.), Compost Science and Technology. Waste Management Series. Elsevier Ltd., pp. 25–48. doi:10.1016/S1478-7482(07)80006-6

Ishii, K., Takii, S., 2003. Comparison of microbial communities in four different composting processes as evaluated by denaturing gradient gel electrophoresis analysis. J. Appl. Microbiol. 95, 109–119. doi:10.1046/j.1365-2672.2003.01949.x

J Jindo, K., Sánchez-Monedero, M.A., Hernández, T., García, C., Furukawa, T., Matsumoto,

K., Sonoki, T., Bastida, F., 2012. Biochar influences the microbial community structure during manure composting with agricultural wastes. Sci. Total Environ. 416, 476–481. doi:10.1016/j.scitotenv.2011.12.009

K Karaca, A., 2010. Biology of earthworms. Springer Science & Business Media.

Kassambara A. 2015. factoextra: Visualization of the outputs of a multivariate analysis [Internet]. R package version 1.0.1. Available: https://cran.r-project.org/web/packages/factoextra/index.html

Kato, K., Miura, N., Tabuchi, H., Nioh, I., 2005. Evaluation of maturity of poultry manure compost by phospholipid fatty acids analysis. Biol. Fertil. Soils 41, 399–410. doi:10.1007/s00374-005-0855-6

Kaushik, P., Garg, V., 2004. Dynamics of biological and chemical parameters during vermicomposting of solid textile mill sludge mixed with cow dung and agricultural residues. Bioresour. Technol. 94, 203–209. doi:10.1016/j.biortech.2003.10.033

Bibliografía

175

Kelessidis, A., Stasinakis, A.S., 2012. Comparative study of the methods used for treatment and final disposal of sewage sludge in European countries. Waste Manage. 32, 1186-1195.

Khwairakpam, M., Bhargava, R., 2009. Vermitechnology for sewage sludge recycling. J. Hazard. Mater. 161, 948–954. doi:10.1016/j.jhazmat.2008.04.088

Killham, K., 1994. Soil ecology. Cambridge University Press.

Kirschbaum, M.U.F., 1995. The temperature dependence of soil organic matter decomposition, and the effect of global warming on soil organic C storage. Soil Biol. Biochem. 27, 753–760. doi:10.1016/0038-0717(94)00242-S

Klamer, M., Bååth, E., 1998. Microbial community dynamics during composting of straw material studied using phospholipid fatty acid analysis. FEMS Microbiol. Ecol. 27, 9–20. doi:10.1016/S0168-6496(98)00051-8

Klammer, S., Knapp, B., Insam, H., Dell’Abate, M.T., Ros, M., 2008. Bacterial community patterns and thermal analyses of composts of various origins. Waste Manag. Res. 26, 173–187. doi:10.1177/0734242X07084113

L Ladd, J.N., Butler, J.H.A., 1972. Short-term assays of soil proteolytic enzyme activities

using proteins and dipeptide derivatives as substrates. Soil Biol. Biochem. 4, 19–30. doi:10.1016/0038-0717(72)90038-7

Lavelle, P., Spain, A. V, 2001. Soil Ecology. Springer, Dordrecht.

Lazcano, C., Gómez-Brandón, M., Domínguez, J., 2008. Comparison of the effectiveness of composting and vermicomposting for the biological stabilization of cattle manure. Chemosphere 72, 1013–1019. doi:10.1016/j.chemosphere.2008.04.016

Leirós, M.., Trasar-Cepeda, C., Seoane, S., Gil-Sotres, F., 1999. Dependence of mineralization of soil organic matter on temperature and moisture. Soil Biol. Biochem. 31, 327–335. doi:10.1016/S0038-0717(98)00129-1

Li, X., Xing, M., Yang, J., Huang, Z., 2011. Compositional and functional features of humic acid-like fractions from vermicomposting of sewage sludge and cow dung. J. Hazard. Mater. 185, 740–8. doi:10.1016/j.jhazmat.2010.09.081

Lin CSK, Pfaltzgraff LA, Herrero-Davila L, Mubofu EB, Abderrahim S, Clark JH, et al. 2013. Food waste as a valuable resource for the production of chemicals, materials and fuels. Current situation and global perspective. Energy Environ Sci.;6: 426–464. doi:10.1039/c2ee23440h

Loehr, R.C., Neuhauser, E.F., Malecki, M.R., 1985. Factors affecting the vermistabilization process. Water Res. 19, 1311–1317. doi:10.1016/0043-1354(85)90187-3

López-González, J.A., Suárez-Estrella, F., Vargas-García, M.C., López, M.J., Jurado, M.M., Moreno, J., 2015. Dynamics of bacterial microbiota during lignocellulosic waste composting: studies upon its structure, functionality and biodiversity. Bioresour. Technol. 175, 406–416. doi:10.1016/j.biortech.2014.10.123

Bibliografía

176

Lores, M., Gómez-Brandón, M., Pérez-Díaz, D., Domínguez, J., 2006. Using FAME profiles for the characterization of animal wastes and vermicomposts. Soil Biol. Biochem. 38, 2993-2996.

Lowe, C.N., Butt, K.R., 2005. Culture techniques for soil dwelling earthworms: A review. Pedobiologia 49, 401–413. doi:10.1016/j.pedobi.2005.04.005

M Madan, R., Pankhurst, C., Hawke, B., Smith, S., 2002. Use of fatty acids for the

identification of AM fungi and the estimation of the biomass of AM spores. Soil Biol. Biochem. 34, 125-128.

MAGRAMA, 2012. Producción y consumo sostenibles y residuos agrarios.

Maulini-Duran, C., Artola, A., Font, X., Sánchez, A., 2014. Gaseous emissions in municipal wastes composting: Effect of the bulking agent. Bioresour. Technol. 172, 260–268. doi:10.1016/j.biortech.2014.09.041

McKinley VL, Vestal JR., 1985.Physical and chemical correlates of microbial activity and biomass in composting municipal sewage sludge. Appl Environ Microbiol.50: 1395–403.

Melgar, R., Benitez, E., Nogales, R., 2009. Bioconversion of wastes from olive oil industries by vermicomposting process using the epigeic earthworm Eisenia andrei. J. Environ. Sci. Health Part B 44, 488–495. doi:10.1080/03601230902935444

Mondini, C., Fornasier, F., Sinicco, T., 2004. Enzymatic activity as a parameter for the characterization of the composting process. Soil Biol. Biochem. 36, 1587–1594. doi:10.1016/j.soilbio.2004.07.008

Monroy, F., Aira, M., Domínguez, J., 2009. Reduction of total coliform numbers during vermicomposting is caused by short-term direct effects of earthworms on microorganisms and depends on the dose of application of pig slurry. Sci. Total Environ. 407, 5411–5416. doi:10.1016/j.scitotenv.2009.06.048

Monroy, F., Aira, M., Domínguez, J., 2008. Changes in density of nematodes, protozoa and total coliforms after transit through the gut of four epigeic earthworms (Oligochaeta). Appl. Soil Ecol. 39, 127–132. doi:10.1016/j.apsoil.2007.11.011

Moreno, J., Moral, R., 2008. Compostaje. Ediciones Mundi-Prensa, Madrid.

Morgan, A.J., 2011. Heavy metals, earthworms, and vermicomposts, in: Edwards, C.A., Arancon, N.Q., Sherman, R. (Eds.), Vermiculture Technology: Earthworms, Organic Wastes, and Environmental Management. CRC Press, Boca Raton, Florida, pp. 263–285.

Mupondi, L.T., Mnkeni, P.N.S., Muchaonyerwa, P., 2010. Effectiveness of combined thermophilic composting and vermicomposting on biodegradation and sanitization of mixtures of dairy manure and waste paper. Afr. J. Biotechnol. 9, 4754–4763. doi:10.5897/AJB09.1894

Bibliografía

177

N Nair, J., Sekiozoic, V., Anda, M., 2006. Effect of pre-composting on vermicomposting of

kitchen waste. Bioresour. Technol. 97, 2091–2095. doi:10.1016/j.biortech.2005.09.020

Nannipieri, P., Kandeler, E., Ruggiero, P., 2002. Enzyme activities and microbiological and biochemical processes in soil, in: Burns, R.G., Dick, R.P. (Eds.), Enzymes in the Environment. New York, pp. 1–34.

Ndegwa, P.M., Thompson, S. A., 2001. Integrating composting and vermicomposting in the treatment and bioconversion of biosolids. Bioresour. Technol. 76, 107–112. doi:10.1016/S0960-8524(00)00104-8

Ndegwa, P.M., Thompson, S. A., Das, K.C., 2000. Effects of stocking density and feeding rate on vermicomposting of biosolids. Bioresour. Technol. 71, 5–12. doi:10.1016/S0960-8524(99)00055-3

Neher, D.A., Weicht, T.R., Bates, S.T., Leff, J.W., Fierer, N., 2013. Changes in Bacterial and Fungal Communities across Compost Recipes, Preparation Methods, and Composting Times. PLoS ONE 8, e79512–e79512. doi:10.1371/journal.pone.0079512

Neuhauser, E.F., Loehr, R.C., Malecki, M.R., 1988. The potential of earthworms for managing sewage sludge, in: Edwards, C.A., Neuhauser, E.F. (Eds.), Earthworms in Waste and Environmental Management. SPB Academic Publishing, the Hague, pp. 9–21.

Nikaeen M, Nafez AH, Bina B, Nabavi BF, Hassanzadeh A., 2015. Respiration and enzymatic activities as indicators of stabilization of sewage sludge composting. Waste Manag. 39: 104–110. doi:10.1016/j.wasman.2015.01.028

Nogales, R., Cifueentes, C., Benitez, E., 2005. Vermicomposting of Winery Waste: A Laboratory Study. J. Environ. Sci. Health B 40, 659–673. doi:10.1081/PFC-200061595

Nolan, T., Troy, S.M., Healy, M.G., Kwapinski, W., Leahy, J.J., Lawlor, P.G., 2011. Characterization of compost produced from separated pig manure and a variety of bulking agents at low initial C/N ratios. Bioresour. Technol. 102, 7131–7138. doi:10.1016/j.biortech.2011.04.066

O Ogunwande GA, Osunade JA, Adekalu KO, Ogunjimi LAO, 2008. Nitrogen loss in chicken

litter compost as affected by carbon to nitrogen ratio and turning frequency. Bioresour Technol.99: 7495–503. doi:10.1016/j.biortech.2008.02.020

Orozco, F., Cegarra, J., Trujillo, L., Roig, A., 1996. Vermicomposting of coffee pulp using the earthworm Eisenia fetida: effects on C and N contents and the availability of nutrients. Biol. Fertil. Soils 22, 162–166. doi:10.1007/BF00384449

Bibliografía

178

P Paradelo, R., Moldes, A.B., Prieto, B., Sandu, R.-G., Barral, M.T., 2010. Can stability and

maturity be evaluated in finished composts from different sources? Compost Sci. Util. 18, 22–31. doi:10.1080/1065657X.2010.10736930

Parthasarathi, K., Ranganathan, L.S., 2000. Aging effect on enzyme activities in pressmud vermicasts of Lampito mauritii (Kinberg) and Eudrilus eugeniae (Kinberg). Biol. Fertil. Soils 30, 347–350. doi:10.1007/s003740050014

Pérez-Losada, M., Eiroa, J., Mato, S., Domínguez, J., 2005. Phylogenetic species delimitation of the earthworms Eisenia fetida (Savigny, 1826) and Eisenia andrei Bouché, 1972 (Oligochaeta, Lumbricidae) based on mitochondrial and nuclear DNA sequences. Pedobiologia 49, 317–324. doi:10.1016/j.pedobi.2005.02.004

Pierzynski, G.M., Vance, G.F., Sims, J.T., 2005. Soils and environmental quality. CRC press.

Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D., R Core Time, 2015. nlme: linear and nonlinear mixed effects models. R package version 3.1-119.

Plaza, C., Nogales, R., Senesi, N., Benitez, E., Polo, A., 2008. Organic matter humification by vermicomposting of cattle manure alone and mixed with two-phase olive pomace. Bioresour. Technol. 99, 5085–5089. doi:10.1016/j.biortech.2007.09.079

Poincelot, R.P., 1972. The Biochemistry and Methodology of Composting. Conn. Agric. Exp. Stn.

Pramanik, P., Chung, Y.R., 2011. Changes in fungal population of fly ash and vinasse mixture during vermicomposting by Eudrilus eugeniae and Eisenia fetida: documentation of cellulase isozymes in vermicompost. Waste Manag. 31, 1169–75. doi:10.1016/j.wasman.2010.12.017

Pramanik, P., Ghosh, G.K., Ghosal, P.K., Banik, P., 2007. Changes in organic - C, N, P and K and enzyme activities in vermicompost of biodegradable organic wastes under liming and microbial inoculants. Bioresour. Technol. 98, 2485–2494. doi:10.1016/j.biortech.2006.09.017

R R Development Core Team, 2014. R: a language and environment for statistical

computing [WWW Document]. R Found. Stat. Comput. Vienna, Austria. URL https://www.r-project.org/

Ranalli, G., Bottura, G., Taddei, P., Garavani, M., Marchetti, R., Sorlini, C., 2001. Composting of solid and sludge residues from agricultural and food industries. Bioindicators of monitoring and compost maturity. J. Environ. Sci. Heal. 36, 415–436. doi:10.1081/ESE-100103473

Bibliografía

179

Ravindran, B., Contreras-Ramos, S.M., Wong, J.W.C., Selvam, A., Sekaran, G., 2014. Nutrient and enzymatic changes of hydrolysed tannery solid waste treated with epigeic earthworm Eudrilus eugeniae and phytotoxicity assessment on selected commercial crops. Environ. Sci. Pollut. Res. 21, 641–651. doi:10.1007/s11356-013-1897-1

Rebollido, R., Martinez, J., Aguilera, Y., Melchor, K., Koerner, I., Stegmann, R., 2008. Microbial populations during composting process of organic fraction of municipal solid waste. Appl. Ecol. Environ. Res. 6, 61–67.

Reinecke, A.J., Reinecke, S. a, Maboeta, M.S., 2001. Cocoon production and viability as endpoints in toxicity testing of heavy metals with three earthworms species. Ecotoxicology 68, 61–68. doi:10.1078/0031-4056-00068

Reynolds, J., Wetzel, M., 2016. Nomenclatura Oligochaetologica–A catalogue of names, descriptions and type specimens. Ill. Nat. Hist. Surv. Spec. Publ.

Richard, T.L., (Bert) Hamelers, H.V.M., Veeken, A., Silva, T., 2002. Moisture Relationships in Composting Processes. Compost Sci. Util. 10, 286–302. doi:10.1080/1065657X.2002.10702093

Richardson, M., 2009. The ecology of the zygomycetes and its impact on environmental exposure. Clin. Microbiol. Infect. 15, 2–9. doi:10.1111/j.1469-0691.2009.02972.x

Riffaldi, R., Levi-Minzi, R., Pera, A., de Bertoldi, M., 1986. Evaluation of compost maturity by means of chemical and microbial analyses. Waste Manag. Res. 4, 387–396. doi:10.1177/0734242X8600400157

Rodríguez-Canché, L.G., Cardoso Vigueros, L., Maldonado-Montiel, T., Martínez-Sanmiguel, M., 2010. Pathogen reduction in septic tank sludge through vermicomposting using Eisenia fetida. Bioresour. Technol. 101, 3548–3553. doi:10.1016/j.biortech.2009.12.001

Romero, E., Plaza, C., Senesi, N., Nogales, R., Polo, A., 2007. Humic acid-like fractions in raw and vermicomposted winery and distillery wastes. Geoderma 139, 397–406. doi:10.1016/j.geoderma.2007.03.009

Ros, M., García, C., Hernández, T., 2006. A full-scale study of treatment of pig slurry by composting: kinetic changes in chemical and microbial properties. Waste Manag. 26, 1108–1118. doi:10.1016/j.wasman.2005.08.008

Ruggieri, L., Artola, A., Gea, T., Sánchez, A., 2008. Biodegradation of animal fats in a co-composting process with wastewater sludge. Int. Biodeterior. Biodegrad. 62, 297–303. doi:10.1016/j.ibiod.2008.02.004

Ruggieri, L., Gea, T., Artola, A., Sánchez, A., 2009. Air filled porosity measurements by air pycnometry in the composting process: a review and a correlation analysis. Bioresour. Technol. 100, 2655–2666. doi:10.1016/j.biortech.2008.12.049

Ryckeboer, J., Mergaert, J., Coosemans, J., Deprins, K., Swings, J., 2003. Microbiological aspects of biowaste during composting in a monitored compost bin. J. Appl. Microbiol. 94, 127–137. doi:10.1046/j.1365-2672.2003.01800.x

Bibliografía

180

Ryckeboer, J., Mergaert, J., Vaes, K., Klammer, S., De Clercq, D., Coosemans, J., Insam, H., Swings, J., 2003. A survey of bacteria and fungi occurring during composting and self-heating processes. Ann. Microbiol. 53, 349–410.

Rynk, R., 2000. Fires At Composting Facilities: Causes And Conditions. BioCycle 41, 54–54.

S Satchell, J.E., Martin, K., 1984. Phosphatase activity in earthworm faeces. Soil Biol.

Biochem. 16, 191–194. doi:10.1016/0038-0717(84)90111-1

Schinner, F., von Mersi, W., 1990. Xylanase-, CM-cellulase- and invertase activity in soil: an improved method. Soil Biol. Biochem. 22, 511–515. doi:10.1016/0038-0717(90)90187-5

Schonholzer, F., Hahn, D., Zeyer, J., 1999. Origins and fate of fungi and bacteria in the gut of Lumbricus terrestris L. studies by image analysis. FEMS Microbiol. Ecol. 28, 235–248.

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. doi:10.1016/j.biortech.2008.07.047

Senesi, N., 1989. Composted materials as organic fertilizers. Adv. Humic Subst. Res. 81, 521–542. doi:10.1016/0048-9697(89)90161-7

Senesi, N., Xing, B., Huang, P.M., 2009. Biophysico-chemical processes involving natural nonliving organic matter in environmental systems. John Wiley & Sons.

Sequi, P., Leita, L., Cercignani, G., 1986. A new index of humification. Agrochimica 30, 175–178.

Shemekite, F., Gómez-Brandón, M., Franke-Whittle, I.H., Praehauser, B., Insam, H., Assefa, F., 2014. Coffee husk composting: an investigation of the process using molecular and non-molecular tools. Waste Manag. 34, 642–652. doi:10.1016/j.wasman.2013.11.010

Sims, G.K., Ellsworth, T.R., Mulvaney, R.L., 1995. Microscale determination of inorganic nitrogen in water and soil extracts. Commun. Soil Sci. Plant Anal. 26, 303–316. doi:10.1080/00103629509369298

Sims, J.R., Haby, V.A., 1971. Simplified colorimetric determination of soil organic matter. Soil Sci. 112, 137–141.

Singh, J., Kalamdhad, A.S., 2013. Effect of Eisenia fetida on speciation of heavy metals during vermicomposting of water hyacinth. Ecol. Eng. 60, 214–223. doi:10.1016/j.ecoleng.2013.07.010

Singh, N.B., Khare, A.K., Bhargava, D., Bhattacharya, S., 2005. Effect of initial substrate pH on vermicomposting using Perionyx excavatus (Perrier, 1872). Appl. Ecol. Environ. Res. 4, 85–97.

Snell-Castro, R., Godon, J.-J., Delgenès, J.-P., Dabert, P., 2005. Characterisation of the microbial diversity in a pig manure storage pit using small subunit rDNA sequence analysis. FEMS Microbiol. Ecol. 52, 229–242. doi:10.1016/j.femsec.2004.11.016

Bibliografía

181

Steger, K., Sjögren, Å.M., Jarvis, Å., Jansson, J.K., Sundh, I., 2007. Development of compost maturity and Actinobacteria populations during full-scale composting of organic household waste. J. Appl. Microbiol. 103, 487–498. doi:10.1111/j.1365-2672.2006.03271.x

Strom, P.F., 1985. Effect of temperature on bacterial species diversity in thermophilic solid-waste composting. Appl. Environ. Microbiol. 50, 899–905.

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, 145–150. doi:10.1016/j.biortech.2004.01.016

Suthar, S., 2010. Pilot-scale vermireactors for sewage sludge stabilization and metal remediation process: comparison with small-scale vermireactors. Ecol. Eng. 36, 703-712.

Suthar, S., 2007. Vermicomposting potential of Perionyx sansibaricus (Perrier) in different waste materials. Bioresour. Technol. 98, 1231–1237. doi:10.1016/j.biortech.2006.05.008

T Thiele-Bruhn, S., 2003. Pharmaceutical antibiotic compounds in soils—a review. J Plant

Nutr Soil Sci 166, 145–67. doi:10.1002/jpln.200390023

Tiquia, S.M., Tam, N.F.Y., 1998. Composting of spent pig litter in turned and forced-aerated piles. Environ. Pollut. 99, 329–337. doi:10.1016/S0269-7491(98)00024-4

Tiquia SM, Tam NFY, Hodgkiss IJ, 1997. Effects of turning frequency on composting of spent pig-manure sawdust litter. Environ Pollut. 62: 37–42. doi:10.1016/S0960-8524(97)00080-1

Tiquia, S.M., Wan, J.H.C., Tam, N.F.Y., 2002. Dynamics of yard trimmings composting as determined by dehydrogenase activity, ATP content, arginine ammonification, and nitrification potential. Process Biochem. 37, 1057–1065. doi:10.1016/S0032-9592(01)00317-X

Tiquia, S.M., Wan, J.H.C., Tam, N.F.Y., 2002. Microbial population dynamics and enzyme activities during composting. Compost Sci. Util. 10, 150–161. doi:10.1080/1065657X.2002.10702075

Tiunov, A.V., Scheu, S., 2004. Carbon availability controls the growth of detritivores (Lumbricidae) and their effect on nitrogen mineralization. Oecologia 138, 83–90. doi:10.1007/s00442-003-1391-4

TMECC, 2002. Test Methods for the Examination of Composting and Compost. Composting Council Research and Education Foundation, and US Department of Agriculture, Bethesda, MD.

Tognetti, C., Mazzarino, M.J., Laos, F., 2007. Improving the quality of municipal organic waste compost. Bioresour. Technol. 98, 1067–1076. doi:10.1016/j.biortech.2006.04.025

Tortosa, G., Castellano-Hinojosa, A., Correa-Galeote, D., Bedmar, E.J., 2017. Evolution of bacterial diversity during two-phase olive mill waste (“alperujo”) composting by 16S rRNA gene pyrosequencing. Bioresour. Technol. 224, 101–111. doi:10.1016/j.biortech.2016.11.098

Bibliografía

182

Trémier, A., Teglia, C., Barrington, S., 2009. Effect of initial physical characteristics on sludge compost performance. Second Int. Conf. Eng. Waste Valoris. 100, 3751–3758. doi:10.1016/j.biortech.2009.01.009

Tuomela M, Vikman M, Hatakka A, Itävaara M., 2000. Biodegradation of lignin in a compost environment: A review. Bioresour Technol. 72: 169–183. doi:10.1016/S0960-8524(99)00104-2

V Vargas-García, M.C., Suárez-Estrella, F., López, M.J., Moreno, J., 2010. Microbial

population dynamics and enzyme activities in composting processes with different starting materials. Waste Manag. 30, 771–778. doi:10.1016/j.wasman.2009.12.019

Ventorino, V., Parillo, R., Testa, A., Viscardi, S., Espresso, F., Pepe, O., 2016. Chestnut green waste composting for sustainable forest management: Microbiota dynamics and impact on plant disease control. J. Environ. Manage. 166, 168–177. doi:10.1016/j.jenvman.2015.10.018

Vig, A.P., Singh, J., Wani, S.H., Singh Dhaliwal, S., 2011. Vermicomposting of tannery sludge mixed with cattle dung into valuable manure using earthworm Eisenia fetida (Savigny). Bioresour. Technol. 102, 7941–7945. doi:10.1016/j.biortech.2011.05.056

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, 1319–26. doi:10.1016/j.biortech.2008.08.014

Voua Otomo, L., Voua Otomo, P., Bezuidenhout, C.C., Maboeta, M.S., 2013. Molecular assessment of commercial and laboratory stocks of Eisenia spp. (Oligochaeta: Lumbricidae) from South Africa. Afr. Invertebr. 54, 499–511. doi:10.5733/afin.054.0220

von Wandruszka, R., 2006. Phosphorus retention in calcareous soils and the effect of organic matter on its mobility. Geochem. Trans. 7, 6. doi:10.1186/1467-4866-7-6

Vuorinen, A.H., 2000. Effect of the bulking agent on acid and alkaline phosphomonoesterase and β-D-glucosidase activities during manure composting. Bioresour. Technol. 75, 133–138. doi:10.1016/S0960-8524(00)00042-0

Vuorinen, A.H., Saharinen, M.H., 1997. Evolution of microbiological and chemical parameters during manure and straw co-composting in a drum composting system. Agric. Ecosyst. Environ. 66, 19–29. doi:10.1016/S0167-8809(97)00069-8

Vu, V.Q., 2011. ggbiplot: A ggplot2 based biplot. R package, version 0.55 [WWW Document]. URL https://github.com/vqv/ggbiplot

Bibliografía

183

w Wagner, M., Loy, A., 2002. Bacterial community composition and function in sewage

treatment systems. Curr. Opin. Biotechnol. 13, 218–227. doi:10.1016/S0958-1669(02)00315-4

Wang, J., Hu, Z., Xu, X., Jiang, X., Zheng, B., Liu, X., Pan, X., Kardol, P., 2014. Emissions of ammonia and greenhouse gases during combined pre-composting and vermicomposting of duck manure. Waste Manag. 34, 1546–1552. doi:10.1016/j.wasman.2014.04.010

Wang, X., Cui, H., Shi, J., Zhao, X., Zhao, Y., Wei, Z., 2015. Relationship between bacterial diversity and environmental parameters during composting of different raw materials. Bioresour. Technol. 198, 395–402. doi:10.1016/j.biortech.2015.09.041

Williams, P.T., 2005. Waste treatment and disposal, second ed. John Wiley & Sons, Chichester.

Williams A.P., Roberts P., Avery L.M., Killham K., Jones D.L., 2006. Earthworms as vectors of Escherichia coli O157:H7 in soil and vermicomposts. FEMS Microbiol. Ecol. 58, 54-64.

Wong, M.H., 1985. Phytotoxicity of refuse compost during the process of maturation. Environ. Pollut. Ser. Ecol. Biol. 37, 159–174. doi:10.1016/0143-1471(85)90006-6

Wong JWC, Mak KF, Chan NW, Lam A, Fang M, Zhou LX, et al, 2001. Co-composting of soybean residues and leaves in Hong Kong. Bioresour Technol.;76: 99–106. doi:10.1016/S0960-8524(00)00103-6

Y Yadav, A., Garg, V.K., 2009. Feasibility of nutrient recovery from industrial sludge by

vermicomposting technology. J. Hazard. Mater. 168, 262–268. doi:10.1016/j.jhazmat.2009.02.035

Yadav, A., Suthar, S., Garg, V.K., 2015. Dynamics of microbiological parameters, enzymatic activities and worm biomass production during vermicomposting of effluent treatment plant sludge of bakery industry. Environ. Sci. Pollut. Res. 22, 14702–14709. doi:10.1007/s11356-015-4672-7

Yadav, K.D., Tare, V., Ahammed, M.M., 2010. Vermicomposting of source-separated human faeces for nutrient recycling. Waste Manage. 30, 50–56.

Yañez, R., Alonso, J.L., Díaz, M.J., 2009. Influence of bulking agent on sewage sludge composting process. Bioresour. Technol. 100, 5827–33. doi:10.1016/j.biortech.2009.05.073

Z Zelles, L., 1999. Fatty acid patterns of phospholipids and lipopolysaccharides in the

characterisation of microbial communities in soil: a review. Biol. Fertil. Soils 29, 111–129. doi:10.1007/s003740050533

Bibliografía

184

Zelles, L., 1997. Phospholipid fatty acid profiles in selected members for soil microbial communities. Chemosphere 35, 275-294.

Zhang, B.G., Li, G.T., Shen, T.S., Wang, J.K., Sun, Z., 2000. Changes in microbial biomass C, N, and P and enzyme activities in soil incubated with the earthworms Metaphire guillelmi or Eisenia fetida. Soil Biol. Biochem. 32, 2055–2062. doi:10.1016/S0038-0717(00)00111-5

Zhang, J., Lv, B., Xing, M., Yang, J., 2015. Tracking the composition and transformation of humic and fulvic acids during vermicomposting of sewage sludge by elemental analysis and fluorescence excitation–emission matrix. Waste Manage. 39, 111-118.

Zhu, W., Yao, W., Du, W., 2016. Heavy metal variation and characterization change of dissolved organic matter (DOM) obtained from composting or vermicomposting pig manure amended with maize straw. Environ. Sci. Pollut. Res. 23, 12128–12139. doi:10.1007/s11356-016-6364-3

Zucconi, F., De Bertoldi, M., 1987. Compost specifications for the production and characterization of compost from municipal solid waste., in: De Bertoldi, M., Ferranti, M.P., L’Hermite, P., Zucconi, F. (Eds.), Production, Quality and Use. Elsevier Applied Science Publisher, London, pp. 30–50.

Zucconi F, Monaco A, Forte M, Bertoldi M., 1985. Phytotoxins during the stabilization of organic matter. In: Gasser JK., editor. Composting of agricultural and other wastes. London: Elsevier Applied Science Publisher. pp. 73–85.

Zucconi, F., Pera, A., Forte, M., de Bertoldi, M., 1981. Evaluating Toxicity of Immature Compost. Biocycle 22, 54–57.