SEMINARIO EL CO2 EN EL CALENTAMIENTO GLOBAL INTRODUCCION Antes de la Era Industrial, cerca 1750, la concentración de dióxido de carbono en la atmósfera (CO2) era de 280 + 10 ppm; pero para 1999 se tenían aproximadamente 367 ppm. Las concentraciones de CO2 en la atmósfera han sido medidas directamente con presición desde 1957 (IPCC, 2001). El incremento de CO2 en la atmósfera en la actualidad es causado por emisiones antropogénicas. Cerca de tres cuartas partes de estas emisiones son debidas a la quema de combustibles fósiles. La quema de combustibles fósiles (más una pequeña contribución de producción de cemento) emitió durante 1980 a 1989 en promedio 5.4 + 0.3 toneladas metricas por año, y 6.3 + 0.4 toneladas metricas por año durante 1990 a 1999. El cambio de uso de la tierra es responsable de el resto de las emisiones ((IPCC, 2001). La concentración de CO2 en la atmósfera es regulada por procesos naturales que intercambian carbono entre la atmósfera, el océano y la biota del suelo. El estudio del ciclo del carbono permite entender y cuantificar los mecanismos, mediante los cuales se transfieren carbono entre y dentro de estos reservorios. Un buen entendimiento del ciclo del carbono es necesario para predecir los niveles de de CO2, y por consiguiente el cambio climático futuro. Los modelos climáticos auguran un importante aumento de la temperatura (1.5 a 4.5°C globalmente) con el doble de CO2. El océano con el tiempo absorverá más del CO2 que se encuentra en exceso en la atmósfera, y el CO2 atmosférico lentamente se estabilizará hacia una concentración asintótica por arriba de los valores pre-industriales. Pero, el alcance de este nuevo equilibrio tomará siglos después de que las emisiones de combustibles fósiles sean detenidas. Las próximas décadas, sin embargo, tanto las plantas como los suelos jugarán un papel muy importante para ayudar al océano a controlar el CO2. Si los árboles en primera instancia toman ventaja en el aumento de CO2 para crecer, la biosfera del suelo ganará carbono y disminuirán los incrementos de CO2 (Rodó y Comín, 2003). El exceso de CO2 atmosférico, debido principalmente a la combustión de combustibles fósiles y un Cambio en el Uso de la Tierra, es un hecho indiscutible. Este continúo aumento de CO2 ha tenido importantes consecuencias para la vegetación. El crecimiento de las plantas es incrementado por excesos de CO2. Un incremento de CO2 puede resultar en un incremento en la producción primaria( Schimel, 1995) y la subsecuente asignación y utilización. Para mitigar las concentraciones de CO2 en la atmósfera se han desarrollado nuevas estratégias y políticas apropiadas para el manejo de la agricultura y los bosques. Una opción se basa en la captura de carbono en los suelos o biomasas terrestres, sobre todo en las tierras usadas para la agricultura o la forestación. A partir del Protocolo de Kyoto esto se conoce como Uso de la Tierra (LU), Cambio en el Uso de la Tierra y Forestación (LULUCF) y concierne a los artículos 1.3 y 1.4 del Protocolo de Kyoto (IPPC, 2000). La toma de acción para la captura de carbono bajo el Protocolo de Kyoto u otro tratado postKyoto no solo estimulará cambios importantes en el manejo del suelo sino que también, por medio de un incremento en el contenido de materia orgánica tendrá efectos significativos directos en sus propiedades y un impacto positivo sobre las cualidades ambientales o agrícolas y sobre la biodiversidad. Las consecuencias incluirán una mayor fertilidad del suelo y productividad de la tierra para la producción de alimentos y para la seguridad alimentaria. Esta herramienta económica también hará que las practicas agrícolas sean más sostenibles y ayudará a prevenir o mitigar la degradación de los recursos de la tierra (FAO, 2001). Este es un trabajo para conocer a grandes rasgos el efecto del CO2 en el calentamiento global, y como afecta este a los microorganismos CARBONO Y DIOXIDO DE CARBONO El carbono es un no metal que se presenta en dos formas naturales: cristalina (diamante y grafito) y amorfa (Hulla, antracita, lignito y turba) y estas variedades alotrópicas se han podido obtener de manera artificial (Alcantara, 1992). El carbono se encuentra libre en la naturaleza y combinado forma varios millones de compuestos. Las sustancias como el petróleo y el gas natural contienen carbono. Todas las plantas y los animales, incluyendo al hombre, contienen carbono en infinita variedad de compuestos (Alcantara, 1992). El carbono es el elemento básico de carbohidratos, grasas, proteínas, ácidos nucleicos (como el DNA y RNA) y otros compuestos orgánicos necesarios para la vida (Miller, 1994). Los principales compuestos del carbono que se convierten en agentes contaminantes son el monóxido de carbono, el dioxido de carbono, el dioxido de azufre, los hidrocarburos, los aldehídos y su interrelaciones quimicas con óxidos de nitrógeno, y otras muchas partículas , por ejemplo, las de fertilizantes y sustancias radiactivas (Alcantara, 1992). Monóxido de Carbono: Es el producto de combustiones incompletas de sustancias orgánicas, las cuales ocurren cuando la cantidad de oxígeno es insuficiente: Compuestos orgánicos + O2 → CO + H2 O 2CO + O2 → CO2 El CO se produce en los escapes de los automóviles, autobuses y camiones, en las emanaciones de combustibles de plantas eléctricas, calentadores, enfriadores, etc. Dioxido de Carbono: Es el producto normal de todas las combustiones de los compuestos orgánicos. CICLO DEL CARBONO El ciclo del carbono está estrechamente ligado al flujo de energía, debido a que las principales reservas de energía de los organismos son compuestos de carbono reducidos que han derivado de la fijación del bióxido de carbono atmosérico ya sea por medio de la fotosíntesis o, con mucho menor frecuencia, de la quimiosíntesis (Miller, 1994). En el Cuadro 1 se se pueden apreciar las cantidades de carbono localizadas en la Tierra: Cuadro 1: Cantidad de carbono presente en varias formas en la tierra Forma en que se halla el carbono Masa de carbono (1018) Carbonato de Ca (La mayor parte en las rocas sedimentarias) 35000 Carbonato de Ca+Mg (la mayor parte en las rocas sedimentarias) 25000 Materia orgánica sedimentaria (como kerógeno) 15000 Carbonato y bicarbonato, disuelto en el océano 42 Combustibles fósiles, recuperables (carbón y petróleo) 4.0 Carbón (humus, caliche, etc) 3.0 Bióxido de carbono atmosférico 0.72 En todos los seres vivos (plantas y animales) 0.56 El ciclo del carbono global: storages El ciclo del carbono en el suelo Los productores primarios en el suelo son principalmente las plantas superirores, aunque las cianobacterias y a un grado menor las algas eucariótas, pueden ser importantes en algunas situaciones como los arrozales o los suelos erosionados y durante la formación del suelo. Por lo general, el bióxido de carbono en la litosfera está a concentraciones mayores que en el aire, pero hay intercambio entre los dos por difusión a través de los espacios del aire del suelo y por medio del equilibrio entre el agua del suelo y el bicarbonato (Campbell, 1987). La materia orgánica llega al suelo a partir de los productores, ya sea de sus hojas, raíces y tallos muertos (conocidos como hojarasca) o de los exudados de las raíces. Se puede considerar que hay tres fuentes principales de materia orgánica en el suelo: insolubles, solubles y microbianas. El carbono insoluble incluye la celulosa, y lignina de las paredes celulares vegetales, quitina de los exoesqueletos de artrópodos y las paredes de algunos hongos, así como otras sustancias que requieren desdoblamiento enzimático antes de que produzcan metabolitos utilizables. El humus es parte del material insoluble. El carbono soluble es aquel que se encuentra en una forma inmediatamente disponible para otros organismos y puede ser liberado por organismos vivos, como las raíces de una planta o después de la descomposición de los productores primarios, consumidores y los mismos microbios degradadores. Una parte considerable del carbón soluble es temporalmente inmovilizado en las células de los microorganismos desintegradotes y su concentración por lo general es baja (Campbell, 1987). El ciclo del carbono en el océano El dióxido de carbono es ràpidamente soluble en agua. Algo de este CO2 disuelto, permanece en el mar, y parte es removido por lo productores fotosintetizantes. Cuanto más caliente esté el agua, mayor es la cantidad de los gases de dióxido de carbono y oxígeno disueltos que regresan a la atmósfera. En los ecosistemas marinos, algunos organismos toman moléculas disueltas de CO2 o iones carbonato (CO3-2) del agua del océano y forman carbonato de calcio ligeramente soluble (CACO3) para elaborar conchas, rocas y esqueleto de los organismos marinos, desde minúsculos protozoarios hasta corales. Cuando los organismos aconchados mueren, partículas finas de sus conchas y huesos caen lentamente a las profundidades del océano, y son enterradas durante eones (o tiempos muy largos) en los sedimentos del fondo. El carbono en estos sedimentos profundos del océano reingresa al ciclo muy lentamente, cuando parte de los sedimentos se disuelvan y formen el dióxido de carbono disuelto que puede entrar a la atmósfera El ciclo del Carbono antes de la Revolución Industrial El océano es con mucho la reserva de carbono más grande cerca de 40000 ton de C. pools over El ciclo del Carbono Después de la Era Industrial Especialmente desde 1950, cuando la población del mundo y el uso de recursos ha aumentado rápidamente, hemos intervenido en el ciclo del carbono principalmente de dos maneras: Eliminación de bosques y otras vegetaciones sin replantación suficiente, lo que deja menos vegetación para absorber CO2. Utilización de combustibles fósiles que contienen carbono y combustión de madera más ràpido de lo que puede volver a reproducirse. Esto produce CO2 que fluye a la atmósfera. Algunos cientificos proyectan que este dióxido de carbono, junto con otras sust Impactos futuros del cambio climático sobre el Ciclo del Carbono En el Océano La cantidad total de carbono en el océano es 50 veces más que la que se encuentra en la atmósfera, y es intercambiada con la atmósfera a gran escala desde hace cientos de años. La respuesta a largo plazo del ciclo del carbono en el océano involucra un cambio climático en las temperaturas de la superficie del mar, en la biología del océano y en la circulación del carbono y nutrientes disueltos. Estos procesos no son independientes uno del otro. Por ejemplo, disminuye la oceanografía de las aguas profundas, rica en nutrientes, y esto puede tener dos consecuncias puestas. Primeramente, se reducirá la presión parcial de CO2 por la de oceanografía de aguas supersaturadas a la superficie, el proceso mediante el cual el agua del mar profundo sube a la superficie upwelling For instance involves changing term response deeper Ecosistemas terrestres EL DIOXIDO DE CARBONO Y EL EFECTO INVERNADERO En un invernadero, el aire se encuentra caliente comparado con el exterior. Este efecto se debe a la energía solar, que en forma de radiaciones ultravioleta visibles, pasa a través deñ vidrio y es absorbida por las plantas y objetos que hay dentro del invernadero. Cuando el interior absorbe la radiación, ésta se convierte en calor; este calor no puede salir a través del vidrio y se concentra en el interior. Esto se conoce como efecto inverndero. Eldióxido de carbono, CO2,actúa de la misma forma que el vidrio de los invernaderos. La energía luminosa y ultravioleta que alcanza la tierra evidentemente es capaz de atravesar la atmósfera, pero después que es absorbida por la tierra y desprendida como energía calorífica, es absorbida por el dióxido de carbono CO2, calentando la atmósfera. Sin este efecto de invernadero, la atmósfera estaría mucho mas fría de que se encuentra. La temperatura promedio de la superficie de la tierra sería mucho menor si no existiera el efecto invernadero. El vapor de agua y las nubes, junto con el dióxido de carbono, contribuyen al efecto invernadero. Sin embargo, como el CO2 es producto de las combustiones, existe la posibilidad de que este efecto se vuelva pernicioso. En efecto, la tala de árboles, el constante aumento de la quema de combustibles fósiles, entre otros factores, alteran la cantidad de dióxido de carbono en l atmósfera, haciendo a ésta más caliente, lo cual, a la larga, tendría efectos nocivos. El sobrecalentamientode la Tierra es una posibilidad real. Y sus efectos serían tan reales como dramaticos. Normalmente, el calor absorbido por el suelo durante el verano se disipa a lo largo del otoño e invierno; po reso las ultimas semanas de invierno y las primeras de primavera suelen ser bastante frías. Sin embargo, el aumento del CO2 atmósferico, al intensificar el efecto invernadero, impide que se disipe todo el calor y que, al verano siguiente, la gente sienta que “hace mas calor que el año pasado”. De este modo va aumentando la cantidad de calor atmosférico, lo cual podría provocar grandes deshielos en las zonas árticas y antárticas, aumentar el nivel de las aguas océanicas e inundar las zonas costeras continentales. Calentamiento Global Adjunto al cambio en las propiedades del litter, los efectos de las concentraciones de los gases de invernadero han sido pronosticadas en los modelos de circulación atmosférica global y hay un consenso general que el efecto primario de un calentamiento global será de 4°C. Por consiguiente ambos aspectos no pueden ser disociados. Agreement Los experimentos de calentamiento de suelo en ecosistemas naturales han servido para generar datos de algunas preguntas importantes incluyendo los efectos del calentamiento del suelo sobre los procesos biquímicos (por ejemplo descomposición, intemperismo del material parental, ciclo del nitrogeno y traza de gas de emisión), la reacción de aquellos cambios en estos procesos puede tener un cambio climático sobre la atmósfera. buried in addition dirigir ardes litter APORTE DE CO2 EN LA ATMOSFERA POR MICROORGANISMOS La función más importante de la flora microbiana es la degradación de materiales orgánicos. El número y diversidad de compuestos disponibles para degradación microbiológica es enorme. El conjunto de ácidos orgánicos, polisacáridos. Ligninas, hidrocarburos alifáticos y aromáticos, azúcares, alcoholes, aminoácidos, purinas, pirimidina, proteínas, lípidos y ácidos nucleicos es atacada por una u otra población. El CO2 que emite el suelo es generado por la respiración de las raíces de las plantas y los organismos, tales como bacterias, hongos, gusanos e insectos (Maddock, 2004). La tasa a la cual el CO2 es liberado durante la mineralización del humus varía ampliamente con el tipo de suelo. Bajo condiciones de laboratorio controladas y a temperaturas en el intervalo mesofílico, 20-30°C, la tasa de producción de CO2 generalmente es de 5 a 50 mg de CO2 por Kg de suelo por día, pero se pueden encontrar en ocasiones 300 mg o más. En el campo, la tasa de formación de CO2 puede ser tan baja como 0.5 o mayor de 10 g de CO2 por metro cuadrado por día y algunas veces se encuentran valores tan altos como 25 (Alexander, 1980). Interacciones entre el enriquecimiento de CO2 atmósférico y la fauna del suelo Es ampliamente aceptado que los incrementos de CO2 en la atmósfera, y otros gases de invernadero, modificaran la fisiología de los ecosistemas. En los ecosistemas terrestres, los efectos de este incremento afectaran al ciclo del carbono, y por ende la traslocación del C dentro de la planta y su transformación por los microorganismos (Madeleine, Bolger, 2000) Montealegre et al., 2002, investigaron la influencia del incremento de CO2 en la atmósfera sobre el número y actividad de las bacterias y las comunidades microbianas del suelo en paztizales bajo atmosferas enriquecidas con dioxido de carbono (FACE). La composición de las comunidades de los microorganismos, en la rizosfera y un volumen de suelo, bajo dos niveles de CO2 bulk free-air Los cambios en la actividad microbiana, número y composición de la comunidad probablemente ocurren bajo elevadas concentraciones de CO2, pero la magnitud de esos cambios va ha depender de la especie de planta y la distancia a la cual los microorganismos esten de las inmediaciones de la las raices de las plantas (Montealegre, 2002). Hay algunas evidencias que los hongos son estimulados más que las bacterias cuando son sometidas a elevadas concentraciones de CO2 (Grayston et al, 1998; Rillig et al., 1999) y diferentes integrantes de las comunidades bacterianas pueden ser afectadas en forma distinta (Marilley et al., 1999; ELhottova et al., 1997; Phillips et al., 2002). Sin embargo, otros estudios no indicaron algún efecto significativo en la estructura de las comunidades microbianas (Zak et al; 2000b) . Muchos reportes presentados, demuestran que a elevadas concentraciones de CO2 se incrementa el porcentaje de infección de las micorrizas vesiculo arbuscular (VAM) como se puede observar en el siguiente cuadro: Cuadro 2: Efecto de elevadas concentraciones de CO2 sobre raices. la Colonización micorrizas de Especie Micorriza Condiciones crecimiento de Porcentaje de infección Referencia Pascopyrum smithii VAM Cárama de crecimiento I Monz et al (1994) Trifolium repens VAM Camara de crecimiento I Jongen et al. (1996) Liriodendron tulipifera VAM Camara abierta de arriba SE O.Neill et al (1991) Beilschmiedia pendula VAM Camara abierta de arriba I Lovelock al. (1996) Pinus taeda EM Invernadero SE Lewis et al (1994) Pinus echinata EM Cárama de crecimiento I O’Neill et al (1987) Pinus radiata EM Cárama de crecimiento SE Conroy et al (1990) Pinus caribaea EM Cárama de crecimiento SE Conroy et al (1990) Pinus silvestres EM Cárama de crecimiento I Ineiche et al (1995) Pinus silvestres EM Cárama de crecimiento SE Perez-Soba et al (1995) Quercus alba EM Cárama de crecimiento I O’Neill et al (1987) Fuente: Luo, Y., y Mooney, H.A. (1999) VAM: Micorriza vesiculoarbuscular ; EM: Ectomicorriza SE:Sin efecto I: Incrementó et El cuadro anterior es una síntesis elaborada por Luo, Y., y Mooney, H.A. (1999) donde mencionan que se tiene que tener cuidado en la interpretación de los datos de porcentaje de infección, ya que solo se usaron datos de porcentajes de infección para evaluar el grado de coliniaciòn micorrizal, pero no la función total del hongo simbiote, la cual es más importante en términos de interacción entre plantas y el componente fúngico. Altos porcentajes de infección no significa una alta cantidad de hifa microrrizal, y viscevesa. Medir la logitud hifal de las micorrizas o biomasa es necesario para resolver este problema. Ronn et al., 2003 midieron el efecto de elevadas concentraciones de CO2 durante el desarrollo del trigo (Triticum aestivum cv. Minaret), y sobre las poblaciones de bacterias y protozoarios del suelo, encontrando que no hubo efecto de la concentración de CO2 en el número de bacterias, pero el número de protozoarios fue alto en los tratamientos con elevadas concentraciones de CO2. Ebersberger et al., 2004 investigaron sobre la estructura de las comunidades microbianas en pastizales de suelos calcareos que han sido expuestos a elevadas concentraciones de CO2 (600 microlitros.L-1) por seis estaciones de crecimiento. Encontrando que en conjunto, solo se detectaron efectos pequeños sobre la estructura de las comunidades microbianas, corroborando previamente que la entrada de carbono al suelo probablemente cambia mucho menos que en la respuesta fotosintética de la planta. La tendencia general en la fauna del suelo respecto a los cambios en el clima son difíciles de predecir porque hay pocos datos y los que existen son generados en experimentos a corto plazo, en areas restringidas y condiciones artificiales. Extrapolaciones a largo plazo y escalas espaciales deben por lo tanto hacerse con precaución porque la fauna del suelo abarca un rango amplio de especies en términos de talla, ciclo de vida, escala de acción, adaptación o potencial de migración que no es bien conocida actualmente, debido a que están interaccionando intimamente con la vegetación y la microflora del suelo. La mayoría de los aspectos del efecto del cambio global sobre la fauna del suelo es pobremente explorada bajo condiciones experimentales y hay un gran rango de prioridades para desarrollar y que puede sobrepasar las especulaciones actuales. Respuesta de las plantas a elevadas concentraciones atmosféricas de CO2. En numerosos estudios se ha reportado que las plantas a elevadas concentraciones de CO2 incrementan su producción de materia seca y algunas veces hacen un uso eficiente del agua (Tubiello et al, 1999). Kang et al, 2002, trabajaron con tres cultivos: trigo, maíz y algodón, dichos cultivos se sometieron a tres contenidos de humedad: alto (85 – 100% de Capacidad de Campo, θF), medio ( 65 – 85% θ) y bajo ( 45 – 65% ); y a dos concentraciones de CO2, baja ( 350 microlitros. L-1) y alta (700 microlitros.l-1), al final de este estudio se llegó a la conclusión que al menos a corto plazo, la plantas C3 tales como el algodón y el trigo pueden beneficiarse de las concentraciones de CO2 especialemente bajo condiciones de escazes de agua • Las elevadas concentraciones de CO2 en la atmósfera probablemente aumentarán la fotosíntesis y crecimiento de las plantas, lo que provocará un incremento en los indices de respiración Sin embargo, en algunos casos los indices de respiración de los tejidos de las plantas se reducen cuando son expuestas a altas concentraciones de CO2, debido a a los efectos directos sobre las enzimas y efectos indirectos derivados de los cambios en la composición química de la planta (Gonzales, 2004) . Gonzales, et al ( 2004) realizaron un revisión bibliográfica sobre la respiración de la planta y elevadas concentraciones de CO2 en la atmósfera y contrario a lo que pensaban, los indices de respiración generalmente no se reducen cuando las plantas se someten a altas concentraciones de CO2. Sin embargo todos los ecosistemas estudiados muestran que la respiración del dosel no se incrementa proporcionalmente al incrementarse la biomasa en respuesta al aumento de la concentración de CO2, aunque una gran proporción de la respiración se lleva a cabo en las raíces. El conocimiento fundamental de cómo la respiración y el soporte de los procesos son controlados fisiológicamente aún es insuficiente, por esta razón se toman medidas acertadas para interpretar como ciertas plantas responden al proceso de respiración cuando se eleva la concentración de CO2. Por lo tanto el papel que juega la respiración de la planta en el aumento de la capacidad de fijacion de los ecosistemas terrestres es aún incierto. La respiración es esencial para el crecimiento y mantenimiento de todos los tejidos de las plantas y juega un papel importante en el balance de carbono en las células, así como también en el ciclo del carbono global. Los ecosistemas terrestres intercambian cerca de 120 toneladas de carbono por año con la atmósfera a través de los procesos de fotosíntesis y respiración (Schlesinger, 1997). Aproximadamente la mitad del CO2 es asimilado anualmente a través de la fotosíntesis y el demás es liberado a la atmósfera por la respiración de las plantas (Gifford, 1994 ; Amthor, 1995 Efecto del CO2 sobre la descomposición de la materia orgánica y el ciclo del nitrógeno Hay una incertidumbre acerca de como los ciclos del carbono y nitrógeno cambiaran con una acumulación de CO2 en la atmósfera. Zak et al., 2000 resumieron datos de 47 publicaciones de los ciclos del carbono y nitrógeno en el suelo bajo elevadas concentraciones de CO2 en un intento por generalizar si los índices disminuirán, incrementaran o no cambiaran. Esta síntesis se centra sobre cambios en la respiración del suelo, respiración microbiana, biomasa microbiana, mineralización total del N, imbolización microbiana y mineralización neta de nitrógeno, porque estas reservas y procesos representan puntos de control importante para el flujo de C y N interno. Determinar si las diferencias en la asignación de C entre las formas de vida de la planta afectan el ciclo del carbono y nitrógeno del suelo de manera predecible, los autores sintetizaron las respuestas en gramineas, herbáceas y árboles que crecen en ambientes con elevadas concentraciones de CO2. Las reservas internas y los procesos que sintetizaron se caracterizaron por un alto grado de variabilidad (coeficiente de variación de 80-800%), hacer generalizaciones dentro y entre formas de vida de la planta es difícil. En pocas excepciones, los indices de respiración microbiana fuero más rápidos bajo concentraciones elevadas de CO2, indicando que 1) generalmente el crecimiento de planta bajo elevadas concentraciones de CO2 incrementa la cantidad de C en el suelo, y 2) un sustrato adicional fue metabolizado por los microorganismos del suelo. Sin embargo, la biomasa microbiana, la mineralización total del N, inmobilización microbiana y la mineralización neta de N son caracterizadas por un gran incremento y disminución bajo concentraciones elevadas de CO2, contribuyendo a un alto grado de varibilidad dentro y entre las formas de vida de planta. De este estudio ellos concluyeron que hay datos insuficientes para predecir como la actividad microbiana y los indices del ciclo del nitrogeno y carbono cambiaran al aumentar las concentraciones de CO2 en la atmósfera.Argumentando que en la actualidad hay lagunas para entender la biología de las raíces finas que limita su habilidad para predecir la respuesta de los microorganismos al incrementar el CO2 en la atmosfera, y entender las diferencias de longevidad de las raíces finas y su bioquímica entre especies de plantas es necesario para desarrollar un modelo que pronostique los cambios en el ciclo del N y C bajo elevadas concentraciones de CO2 atmosférico. El cambio climático afectará directamente la mineralización del carbono y nitrogeno a través de los cambios de temperatura y humedad del suelo, pero indirectamente puede afectar los indices de mineralización a través de los cambios en la calidad del suelo. Keller et al (2004) realizaron un experimento en un mesocosmo para examinar los efectos de seis años de rayos infrarrojos (calentamiento) y láminas de agua sobre el potencial anaerobico del nitrogeno y carbono water-table donde in bog fen peat frared loading pathways La predicción de que la calidad del matillo, y por lo tanto los indices de descomposición del mantillo, pueden reducir cuando las plantas crecen en atmósferas enriquecidas con CO2 ha sido basada en la observación de la concentración de N foliar que generalmente es baja en elevadas concentraciones de CO2. Norby, et al (2000) evaluaron el supuesto que el proceso de resorción de nutriente estacional es diferente en plantas eriquecidas con CO2. La resorción del N fue estudiada en dos especies de arboles de maple (Acer rubrum L y A. saccharum Marsh). Estos encontraron que los efectos de elevadas concentraciones de CO2 sobre el N del litter son intrínsecamente más difíciles de detectar que las diferencias en hojas verdes porque los factores que afectan la senescencia y resorción son muy variables. Los resultados de este experimento soportan la aproximación usada en ecosistemas modelo en los cuales la efectividad de resorción es constante a ambiente y elevadas concentraciones de CO2, pero los resultados tambien indican que otros factores pueden alterar la eficiencia de resorción. La descomposición del mantillo en bosques Mediterraneos disminuye bajo elevadas concentraciones de CO2, pero si este efecto es combinado a un aumento de producción primaria , aumentará el almacenamiento de C en suelos de ecosistemas forestales. Entonces los suelos forestales, por consiguiente, representan un sumidero potencial para los excesos de carbono (Angelis et al., 2000). Torbert et al., (2000) utilizaron cultivos de algodón, trigo, sorgo y soya para analizar los impactos de elevadas concentraciones de CO2 sobre la descomposición de los residuos de estos cultivos dentro de agroecosistemas, Las observaciones de campo y laboratorio de este estudio indican que con elevadas atmosferas de CO2, el indice de descomposición de los residuos de la planta puede ser limitado por el N y la liberación de N de la descomposición de los residuos de la planta pueden ser bajos. ALTERNATIVAS PARA MITIFAR CO2 EN LA ATMOSFERA CAPTURA DE CARBONO La fotosíntesis remueve alrededor de 105 Pg C (1 Pg=1*109) de la atmósfera cada año, cerca de la mitad es tomado por la plantas, releases with just over order removes. Las plantas de la biosfera pueden tener un enorme impacto sobre los niveles de CO2 atmosféricos Revisar artículo . Actualmente, la producción de cultivos produce una modesta cantidad de CO2 a la atmósfera (AAFC, 1999). Con un manejo correcto y un cambio de cultivo que incremente la cantidad de materia orgánica y reduzca la los indices de descomposición del material orgánico del suelo, los suelos agrícolas pueden llegar a ser el mejor sumidero de CO2 sink crop change Currently. Captura de Carbono en tierras áridas Las tierras áridas se definen por en índice de aridez que representa la realación de la precipitación con la evapotranspiración potencial (P/PET) con valores <0.05 para tierras hiperáridas, <0.20 para tierras áridas y de 0.20 a 0.50 para tierras semiáridas. Estas son las tierras secas más características, pero a menudo la zona árida subhúmeda (0.5-0.65) tambien se incluye en la misma (Middleton y Thomas, 1997) citado por FAO ( 2001). Las tierras áridas representan cerca del 40% de las tierras del globo. Las zonas hiperáridas naturales cubren un área estimada en 1000 millones de hectáreas mientras que las tierras áridas, semiáridas y áridas subhumedas cubren un área de 5100 millones de hectáreas. Si bien el contenido de carbono y la capacidad de fijar CO2 por unidad de superficie en las tierras áridas son bajos, pueden de cualquier manera hacer una contribución importante a la captura global de carbono y al mismo tiempo prevenir o diminuir la tasa de desertificación. Con esta amplia definición, una gran parte de las tierras áridas se incluyen en el área tropical definida como la parte intertropical del mundo, la que representa el 37.2 por ciento de la superficie terrestre (4900 millones de hectáreas) En un estudio realizado en el desierto de Negev en Yatir, USA con el pino Aleppo (Pinus halepensis Mill) durante 35 años encontraron que el aumento de CO2 en la atmósfera puede conducir a un aumento en la disponibilidad del agua, debido a un uso eficiente del agua por las plantas al incrementar el CO2. Esto permite extender las actividades forestales en regiones aridas. Estas 2800 ha de tierras forestales aridas contiene 6.5+ 1.2 kg Cm-2 y continúa acumulandose en 0.13-0.24 kg Cm-2 año-1. El CO2 es alto durante el invierno, Los esfuerzos de reforestación dentro de regiones áridas puede ser significativo para la captura de C y los beneficios (restauración de la tierra degradada, reducción de ecorrentía, erosión y compactación del suelo, beneficio a la vida salvaje) debido a la gran extensión de las regiones potencialmente involucradas ( aproximadamente 2*109 de las tierras con arbustos y paztizales C4). Información cuantitativa de las actividades forestales bajo condiciones áridas es relevante para regiones que preveen un aumento en la aridez (Grunzweig et al., 2003 ). Captura de Carbono en áreas tropicales Captura de carbono en tierras de pastoreo Las tierras de pastoreo juegan un papel muy importante en la captura de carbono. En primer lugar, las tierras de pastoreo, según FAO, ocupan 3200 millones de hectáreas y almacenan entre 200 y 420 Pg en el ecosistema total, una gran parte del mismo debajo de la superficie y, por lo tanto, en un estado relativamente estable. El carbono del suelo en las tierras de pastoreo es estimado en 70 t/ha, cifra similar a las cantidades alamacenadas en los suelos forestales (Trumbmore et al., 1995; Balesdent y Arrouays, 1999) citado por (FAO, 2001). Muchas áreas de tierras de pastoreo en la zonas tropicales y àridas son mal manejadas y están degradadas; por lo tanto ofrecen variadas posibilidades de suecuestro de carbono. Captura de carbono en ecosistemas forestales Los bosques cubren el 29% de las tierras y contienen el 60 % del carbono de la vegetación terrestre. El carbono almacenado en los suelos forestales representa el 36% del total del carbono del suelo a un metro de profundidad (1500 Pg). Los ecosistemas forestale contienen más carbono por unidad de superficie que cualquier otro tipo de uso de la tierra y sus suelos, que contienen cerca del 40% del total del carbono son de importancia primaria cuando se considera el manejo de los bosques. Por lo general, en los bosques naturales el carbono del suelo está en equilibrio, pero tan pronto como ocurre la deforestación o la reforestación, ese equilibrio es afectado. En los lugares donde no se puede ser detenida la deforestación, es necesario un manejo correcto para minimizar las pérdidas por carbono. La reforestación, sobre todo en los suelos degradados con bajo contenido de materia orgánica, será una forma importante de secuestro de carbono a largo plazo, tanto en la biomasa como en el suelo. Comparado con los demás ecosistemas, los suelos turbosos son un pequeño sumidero de dioxido de carbono (Turunen et al., 2002). Captura de carbono en tierras cultivadas En los sistemas agrícolas, una gran parte de carbono es almacenado en el suelo. La entrada de carbono en los suelos es determinada por la producción primaria neta y la fracción que permanece sobre el campo. La pérdida de carbono está determinada por la descomposición y por la erosión de la capa arable. La descomposición es controlada por la temperatura ambiente y las propiedades fisicas y químicas del suelo. En general, los rendimientos de los cultivos son bajos, los contenidos de carbono en el suelo son altos y los indices de descomposición de la materia orgánica en el suelo son altos lo cual incrementa la pérdida de carbono (Freibauer, 2004). outside Enhance focus overall sink baseline El desarrollo de la agricultura ha implicado una gran pérdida de materia orgánica del suelo. Hay varias formas de las diferentes prácticas de manejo de tierras que pueden ser usadas para aumentar el contenido de materia orgánica del suelo, tales como el incremento de la productividad y de la biomasa –variedades, fertilización e irrigación. El cambio climático global puede tener un efecto similar. Las principales formas de obtener un incremento de la materia orgánica en el suelo están asociadas a la agricultura de conservación y conforman la labranza mínima o cero y el uso de cobertura vegetal continúa y protectora formada por materiales vegetales vivos o muertos sobre la superficie del suelo. Baseline whether depends sink lead PRINCIPALES CONSECUENCIAS E IMPACTO DE LA CAPTURA DE CARBONO La captura de carbono y el aumento de la materia orgánica del suelo tendrán un impacto directo sobre la calidad y la fertilidad de los suelos. Habrá también efectos positivos importantes sobre el ambiente y la resiliencia y la sostenibilidad de la agricultura. Calidad y fertilidad del suelo La materia orgánica del suelo tiene funciones esenciales desde el punto de vista biológico, físico y químico del suelo. El contenido de materia orgánica es generalmente considerado como uno de los indicadores primarios de la calidad del suelo, tanto en sus funciones agrícolas como ambientales. La materia orgánica es de especial interés en el caso de los suelos tropicales -excepto en los vertisoles- con arcillas de baja actividad que tienen una pobre capacidad de intercambio de cationes. La capacidad de intercambio de cationes aumenta en función del incremento de la materia orgánica . La biodisponibilidad de otros elementos importantes tales como el fósforo podrá mejorar y la toxicidad de otros elementos podrá ser inhibida por la formación de quelatos u otras uniones, por ejemplo, aluminio y materia orgánica (Robert, 1996a). En una agricultura con bajo uso de insumos, el reciclaje de los nutrimentos -N, P, K, Ca- por medio de la descomposición gradual de las plantas y los residuos de los cultivos es de importancia fundamental para la sostenibilidad (Sánchez y Salinas, 1982; Poss, 1991). En el caso de la erosión, se ha establecido una correlación entre la disminución histórica de la materia orgánica del suelo y el desarrollo de la erosión. Todos los tipos de manejo de los cultivos que capturan carbono favorecen la cobertura del suelo y limitan la labranza y de este modo previenen la erosión. Impactos ambientales La captura de carbono en los suelos agrícolas se contrapone al proceso de desertificación por medio del papel que juega el incremento de la materia orgánica sobre la estabilidad de la estructura -resistencia a la erosión hídrica y eólica- y a la retención de agua, y al aspecto esencial de la cobertura de la superficie del suelo directamente por las plantas o por los residuos de las plantas -o cobertura muerta- para prevenir la erosión e incrementar la conservación del agua (FAO, 2001). La materia orgánica, al incrementar la calidad del suelo, también tiene una función protectiva al fijar los contaminantes -ya sean orgánicos como los pesticidas o minerales como los metales pesados o el aluminio- los cuales, en general, disminuyen en su toxicidad (FAO, 2001). La calidad del aire está principalmente relacionada con la disminución de la concentración del CO2 atmosférico, pero considerando también los otros gases de invernadero, en particular metano y óxido nitroso (CH4 y N2O). El principal factor que controla su génesis es la anaerobiosis proceso de reducción del suelo- la cual está generalmente ligada a las condiciones hidromórficas. Cuando aumentan las pasturas o las tierras para pastoreo, la emisión de metano por el ganado debe también ser tomada en consideración (FAO, 2001). En algunos ambientes y dependiendo de las condiciones climáticas -áreas húmedas- o propiedades del suelo -alto contenido de arcilla- puede ser formado N2O. Por lo tanto, se debe hacer un cuidadoso balance de los distintas emisiones de gases. El cultivo del arroz en tierras húmedas es el sistema de cultivos más complejo en relación a la captura de carbono. Si la materia orgánica se acumula en un suelo húmedo, también se forma CH4. El efecto de invernadero del metano es mucho mayor que el del CO2. La estrategia más común para prevenir la formación de metano es disminuir el período de inundación, de modo que la materia orgánica esté menos protegida de la mineralización y puedan ser emitidos CO2 y N2O o NH4. Por estas razones, parecería muy difícil, por el momento, manejar simultáneamente la producción de arroz en tierras húmedas y la captura de carbono. Los últimos hallazgos en la agricultura de conservación respecto a los sistemas arroz-trigo son positivos; por ejemplo, los rendimientos del arroz pueden ser mantenidos o mejorados sin saturación de agua, encharcamiento o reducción del suelo y con grandes ahorros de agua en el período de crecimiento del arroz. Este nuevo enfoque ha sido convalidado por los agricultores en varios miles de hectáreas en países como India y Brasil. Las tierras húmedas tienen condiciones anaeróbicas similares con menor emisión de CH4 que los campos de arroz húmedos y un mayor potencial de secuestro de carbono que puede llevar a la formación de turba. Esto tiene también otras ventajas ambientales importantes que deben ser protegidas; no es realista, sin embargo, esperar rápidos incrementos. La calidad del agua también es mejorada por una disminución de la escorrentía, de los contaminantes y de la erosión. En el caso específico de la labranza de conservación, también se evita o minimiza una fuerte mineralización de la materia orgánica con la subsecuente formación de nitratos. Los cambios en el uso de la tierra y en su manejo también tienen un efecto importante sobre la distribución de la precipitación pluvial entre escorrentía y almacenamiento o infiltración, con un aumento de la última en el caso de las tierras de pastoreo, bosques y labranza de conservación con cobertura de suelo. La cobertura del suelo previene la erosión; por lo tanto, si hubiera alguna escorrentía, el agua estaría libre de partículas asociadas con contaminantes -elementos traza, PO4. La contaminación a distancia por productos solubles también disminuirá en relación directa con la menor escorrentía. Esta es una de las bases de la ecocondicionalidad en la US Farm Bill desde 1996. Con tales cambios en las prácticas agrícolas puede ser enfrentado el desafío de la calidad del agua. Una vez que los cambios hayan tenido lugar en grandes áreas, también decrecerá la severidad y frecuencia de las inundaciones. El efecto general del incremento de la materia orgánica del suelo es un mejoramiento de la capacidad amortiguadora y de la resiliencia del suelo a diferentes tipos de degradación o estrés. Biodiversidad y función biológica del suelo Los cambios en la biodiversidad son evidentes cuando ocurre la deforestación. En el caso de la reforestación, dependerán del tipo de bosque establecido. Los sistemas agro-forestales bien manejados involucran una amplia gama de biodiversidad. Por lo general, la biodiversidad de los mamíferos es preservada en el caso de los bosques, el número de especies de aves se reduce a la mitad y las especies vegetales disminuyen en un tercio (de 420 a 300), (IPCC, 2000). Likey, (ICRAF) se refiere a un mosaico de manchas, cada una de ellas compuesta de muchos nichos, o sea un sistema favorable para la biodiversidad. En el pasado, los sistemas agrícolas más intensivos llevaron a una sensible disminución de la biodiversidad, junto a una paralela reducción de la materia orgánica debida sobre todo a la labranza y al uso de pesticidas (Rovira, 1994). En el caso de las tierras de cultivo, el aumento de la biodiversidad en relación con el incremento de la materia orgánica se basa, sobre todo, en el aumento de la biodiversidad del suelo (Copley, 2000). La Figura 13 presenta una organización jerárquica de la biodiversidad del suelo, la cual depende directamente del abastecimiento de materia orgánica fresca y de las prácticas agronómicas. Esta biodiversidad varía desde los genes hasta los microorganismos, la fauna y la biodiversidad encima de la tierra. La cantidad de bacterias puede aumentar en forma exponencial, de 103 a 1012, tan pronto como la materia orgánica sea abundante. La labranza cero favorecerá el desarrollo de hongos los cuales son sumamente activos en la agregación del suelo. Sin embargo, solo 5 a 10 por ciento de las especies de la microflora del suelo son conocidas y en la actualidad sería posible investigar, gracias a las nuevas técnicas moleculares, la evaluación de la biodiversidad específica o interespecífica de los microorganismos. Figura 13. Organización jerárquica de la biodiversidad del suelo Cuando la materia orgánica fresca -residuos de las plantas o plantas de cobertura- está presente en la superficie del suelo, habrá un incremento de las distintas categorías de la fauna, sobre todo de los descomponedores. Las cadenas alimenticias asociadas a los detritos serán estimuladas (Hendricks et al., 1986) -bacterias, hongos, microartró-podos, nematodos, enquitreidosmacroartrópodos. Las lombrices de tierra, las termites y las hormigas, que son los principales grupos que componen la macrofauna (>1 cm) a menudo son llamados ingenieros del suelo en razón de la función que tienen sobre la porosidad -bioporos- y estructura del suelo; su número se incrementa paralelamente al aumento de la materia orgánica con una disminución del disturbio del suelo, o sea la no labranza (Figura 14). Son buenos indicadores de la calidad del suelo (Lavelle, 2000; Lobry de Bruyn, 1997) y tienen un papel fundamental en la agricultura de conservación. Son, por ejemplo, indispensables para asegurar la distribución a través del suelo -incluso a más de un metro de profundidad- de la materia orgánica acumulada en la superficie. Figura 14. Efecto del sistema anterior de labranza sobre el número de lombrices de tierra en varias fincas. Canterbury, Nueva Zelandia (de Fraser, en Soil biota, 1994) Un aumento en la captura de carbono causa un incremento en la biodiversidad activa y un funcionamiento más efectivo de los elementos biológicos del suelo, lo cual es un proceso relativamente lento en la mayoría de los suelos agrícolas. La biodiversidad de todo el agro-sistema (vegetación, aves, etc.) también depende del tipo de manejo. Todas las consecuencias y los beneficios de este enfoque también deberían ser apreciados en relación con la sostenibilidad de la agricultura, incluso con respecto a los depósitos de genes y el control biológico de las plagas. Efectos del cambio climático Mientras que un aumento del contenido atmosférico de gases de invernadero está llevando a un cambio climático, también ocurrirán numerosos efectos complejos, contrastantes y opuestos (Brinkman y Sombroek, 1996). Todos los resultados experimentales demuestran que un aumento de la concentración de CO2 en la atmósfera induce un incremento de la biomasa o de la Red Primaria de Producción por medio de la fertilización con carbono, con un papel muy importante sobre la fotosíntesis y el crecimiento de las plantas. La ganancia en la fijación de CO2 podría ser importante. El incremento en la productividad medido a causa de la duplicación de la concentración del CO2 -predicha para el año 2100- es de cerca del 30 por ciento para las plantas C3. Otro efecto importante del aumento del CO2 es la disminución de la transpiración de las plantas a través de los estomas lo cual redunda en una mayor eficiencia en el uso del agua (WUE), sobre todo en las plantas C4. En lo que se refiere al agua, hay un efecto neto favorable del CO2 sobre la reducción de la transpiración de las plantas (Gregory et al., 1998). Evidentemente, para llegar a un aumento de rendimiento en el campo, también deben ser satisfechos otros requerimientos de las plantas como el agua y los nutrientes disponibles. En lo que se refiere al ciclo del carbono, habrá una mayor captura de carbono por la biomasa aérea y un correlativo ingreso de carbono en el suelo a partir de los residuos de las plantas y del crecimiento y la muerte de las raíces más finas. Los compuestos de las raíces tienen una mayor relación C/N y son más estables. Otro factor que juega un papel importante en la captura de carbono es la temperatura, la que podría aumentar en algunas partes del globo terráqueo. Tal incremento podría provocar una mayor tasa de mineralización de la materia orgánica por los microorganismos y una mayor tasa de respiración de las raíces. Este efecto de la temperatura sobre la mineralización podría ser significativo en los países fríos, donde la temperatura es un factor limitante y donde puede ser esperado un incremento de las emisiones de CO2. Sin embargo, en la mayor parte del mundo es de esperar un aumento de la captura de carbono (van Ginkel et al., 1999). Para estimar el efecto del cambio climático sobre la captura de carbono pueden ser usados modelos. Los resultados de muchos estudios recientes confirman el incremento de la tasa de crecimiento de los bosques en las zonas templadas y en los países nórdicos. En lo que se refiere a los bosques tropicales, existen algunas medidas hechas en la Amazonía donde se ha encontrado un aumento de la biomasa (Phillips et al., 1998) de 0,62 t C/ha/año, lo cual para un área de 7 000 millones de hectáreas significa una captura de carbono de Gt 0,44 C/año. La causa de esto no es simple, ya que la influencia de El Niño está probablemente involucrada en el aumento de la humedad del área. USO DE COMBUSTIBLES O ENERGIA ALTERNATIVA CONCLUSIONES BIBLIOGRAFIA Angelis, P., Chigwerewe, K.S. and Scarascia Mugnozza, G.E. 2000. Litter quality and decomposition in a CO2-enriched Mediterranean forest ecosystem. Plant and Soil 224: 31-41 Luo, Y. and Mooney, H.A. 1999. Carbon Dioxide and Enviromental Stress. Ed. Academic Press, San Diego California, USA. . 418 pp. Norby, R.J., Long, T.M., Hartz-Rubin, J.S and O’Neill, E.G. 2000. Nitrogen resorption in senescing tree leaves in a warmer, CO2-enriched atmosphere. Plant and Soil 224: 15-29. Torbert, H.A., Prior, S.A., Rogers, H.H. and Wood, C.W. 2000. Review of elevated atmospheric CO2 effects on agro-ecosystems:residue decomposition processes and soil C storage. Plant and Soil 224: 59-73. DIAGRAMA DEL PROCESO DE PRODUCCION DE SETAS Y RECICLAJE DE LA PAJA DE CEBADA, PROVENIENTE DE LA PRODUCCION DEL HONGO SETA (Pleurotos ostreatus) Remojado de pacas de cebada Pasteurización (45 min.) Escurrimiento y encalado (40 min) Siembra Incubación (20-40 días) Cosecha (60- 90 días) Residuo de paja de cebada de tres oleadas Residuo de paja de cebada de cinco oleadas oleadas Residuo de paja de cebada contaminada Mezcla de pajas Elaboración de pilas y llenado de las pilas Inoculación de lombriz de tierra Control de temperatura, aireación y humedad de la vermicomposta Caracterización del compost ANALISIS DE GRUPOS MICROBIANOS EN EL PROCESO DE PRODUCCION DE COMPOSTA Y VERMICOMPOSTA Erica Morales Hernández1 1 Alumna de la Especialidad de Edafología, IRENAT-Colegio de Postgraduados, Montecillos, Estado de México La cinética de los grupos microbianos es muy importante en la producción de compost y vermicompost, ya que éstos junto con los factores climáticos y la humedad pueden disminuir el tiempo para degradar la materia orgánica a compuestos más estables y asimilables por las plantas. Por otra parte la incorporación de lombrices al proceso de composteo ayuda y favorece en gran medida a la transformación de los residuos orgánicos sin valor comercial a productos con valor comercial. Es recomendable realizar un análisis fisicoquímico y microbiológico cuando un residuo ha sido previamente sometido a algún tratamiento, ya sea físico o químico, antes de empezar con el proceso de composteo y vermicomposteo y aun más cuando los residuos han recibido un precomposteo, como es el caso del sustrato que es utilizado para el crecimiento del hongo Pleurotos ostreatus y residuos hortofrutícolas , ya que en el primero, la paja de avena ha recibido una pasteurización, el hongo ha consumido lignina y celulosa principalmente, y además se quedan residuos de micelio y cuerpos fructíferos que tienen un alto contenido de proteínas, lo que da lugar a que el residuo sea más rápidamente degradado, mientras que en el segundo, los microorganismos ya han degrado moléculas como almidón y ha aumentado la población de microorganismos, lo que favorece el proceso de vermicomposteo y composteo. Además es recomendable realizar análisis fisicoquímicos y microbiológicos del compost y vermicompost; dentro del químico recomendaría iones solubles (calcio, magnesio, sodio, potasio, cloruros, carbonatos, bicarbonatos, sulfatos, nitratos, amonio, fosfatos, boro, molibdeno), ya que estos son fácilmente disponibles para las plantas, y con estos datos podría hacer una balance nutrimental de acuerdo al cultivo que se desee manejar, ya que los análisis que le hicieron de Nitrógeno y Fósforo total al compost y vermicompost en este trabajo, no se sabe si están disponible o como macromoléculas todavía no asimilables por las plantas, y la conductividad eléctrica solo es un indicador de salinidad, pero no nutrimental, y lo que realmente nos interesa cuando se realiza una composta es su calidad nutrimental, ya que de eso depende si se utiliza o no con fines agrícolas. En cuanto a los análisis físicos recomendaría humedad, curva de tensión de humedad, porosidad, infiltración, ya que esto es necesario cuando se usa como sustrato para la producción de algún cultivo en particular. BIBLIOGRAFIA Corlay, C., L., Ferrera, C., R., Etcheverts, B., J., Echegaray, A., Alfredo y Santizo, R., J. 1999. Cinetica de grupos microbianos en el proceso de producción de composta y vermicomposta. Agrociencia 33, 375-380.. New Volume 147 Issue doi:10.1046/j.1469-8137.2000.00687.x 1 Page 201 Elevated atmospheric CO2, fine roots microorganisms: a review and hypothesis - and the July response Phytologist 2000 of DONALD R. ZAK, KURT S. PREGITZER, JOHN S. KING & WILLIAM E. HOLMES There is considerable uncertainty about how rates of soil carbon © and nitrogen (N) cycling will change as CO2 accumulates in the Earth’s atmosphere. We summarized data from 47 published reports on soil C and N cycling under elevated CO2 in an attempt to generalize whether rates will increase, decrease, or not change. Our synthesis centres on changes in soil respiration, microbial respiration, microbial biomass, gross N mineralization, microbial immobilization and net N mineralization, because these pools and processes represent important control points for the belowground flow of C and N. To determine whether differences in C allocation between plant life forms influence soil C and N cycling in a predictable manner, we summarized responses beneath graminoid, herbaceous and woody plants grown under ambient and elevated atmospheric CO 2. The belowground pools and processes that we summarized are characterized by a high degree of variability (coefficient of variation 80 800%), making generalizations within and between plant life forms difficult. With few exceptions, rates of soil and microbial respiration were more rapid under elevated CO 2, indicating that (1) greater plant growth under elevated CO 2 enhanced the amount of C entering the soil, and (2) additional substrate was being metabolized by soil microorganisms. However, microbial biomass, gross N mineralization, microbial immobilization and net N mineralization are characterized by large increases and declines under elevated CO2, contributing to a high degree of variability within and between plant life forms. From this analysis we conclude that there are insufficient data to predict how microbial activity and rates of soil C and N cycling will change as the atmospheric CO 2 concentration continues to rise. We argue that current gaps in our understanding of fine-root biology limit our ability to predict the response of soil microorganisms to rising atmospheric CO 2, and that understanding differences in fine-root longevity and biochemistry between plant species are necessary for developing a predictive model of soil C and N cycling under elevated CO 2. soil Forward Links to Citing Articles • Veronika ezá ová, Herbert Blum, Hana Hr elová, Hannes Gamper and Milan Gryndler. (2005) Saprobic microfungi under Lolium perenne and Trifolium repens at different fertilization intensities and elevated atmospheric CO2 concentration. Global Change Biology 11:2, 224-230 • S. A. Billings and S. E. Ziegler. (2005) Linking microbial activity and soil organic matter transformations in forest soils under elevated CO2. Global Change Biology 11:2, 203-212 • Anne Kasurinen, Paula Kokko-Gonzales, Johanna Riikonen, Elina Vapaavuori and Toini Holopainen. (2004) Soil CO2 efflux of two silver birch clones exposed to elevated CO2 and O3 levels during three growing seasons. Global Change Biology 10:10, 1654-1665 • Sini Maaria Niinistö, Jouko Silvola and Seppo Kellomäki. (2004) Soil CO2 efflux in a boreal pine forest under atmospheric CO2 enrichment and air warming. Global Change Biology 10:8, 1363-1376 • John S. King, Paul J. Hanson, Emily Bernhardt, Paolo DeAngelis, Richard J. Norby and Kurt S. Pregitzer. (2004) A multiyear synthesis of soil respiration responses to elevated atmospheric CO2 from four forest FACE experiments. Global Change Biology 10:6, 1027-1042 • Elise Pendall, Scott Bridgham, Paul J. Hanson, Bruce Hungate, David W. Kicklighter, Dale W. Johnson, Beverly E. Law, Yiqi Luo, J. Patrick Megonigal, Maria Olsrud, Michael G. Ryan and Shiqiang Wan. (2004) Below-ground process responses to elevated CO2 and temperature: a discussion of observations, measurement methods, and models. New Phytologist 162:2, 311-322 • Robert S. Nowak, David S. Ellsworth and Stanley D. Smith. (2004) Functional responses of plants to elevated atmospheric CO2 do photosynthetic and productivity data from FACE experiments support early predictions?. New Phytologist 162:2, 253-280 • Shiqiang Wan, Richard J. Norby, Kurt S. Pregitzer, Joanne Ledford and Elizabeth G. O’Neill. (2004) CO2 enrichment and warming of the atmosphere enhance both productivity and mortality of maple tree fine roots. New Phytologist 162:2, 437-446 • Romain Barnard, Laure Barthes, Xavier Le Roux and Paul W. Leadley . (2004) Dynamics of nitrifying activities, denitrifying activities and nitrogen in grassland mesocosms as altered by elevated CO2. New Phytologist 162:2, 365-376 • William E. Holmes, Donald R. Zak, Kurt S. Pregitzer and John S. King. (2003) Soil nitrogen transformations under Populus tremuloides, Betula papyrifera and Acer saccharum following 3 years exposure to elevated CO2 and O3. Global Change Biology 9:12, 1743-1750 • Graham J. Hymus, David P. Johnson, Sabina Dore, Hans P. Anderson, C. Ross Hinkle and Bert G. Drake. (2003) Effects of elevated atmospheric CO2 on net ecosystem CO2 exchange of a scruboak ecosystem. Global Change Biology 9:12, 1802-1812 • Hormoz BassiriRad, John V. H. Constable, John Lussenhop, Bruce A. Kimball, Richard J. Norby, Walter C. Oechel, Peter B. Reich, William H. Schlesinger, Stephen Zitzer, Harbans L. Sehtiya, and Salim Silim. (2003) Widespread foliage 15N depletion under elevated CO2: inferences for the nitrogen cycle. Global Change Biology 9:11, 1582-1590 • Karina V. R. Schäfer, Ram Oren, David S. Ellsworth, Chun-Ta Lai, Jeffrey D. Herrick, Adrien C. Finzi, Daniel D. Richter, and Gabriel G. Katul. (2003) Exposure to an enriched CO2 atmosphere alters carbon assimilation and allocation in a pine forest ecosystem. Global Change Biology 9:10, 1378-1400 • JOHN R. BUTNOR, KURT H. JOHNSEN, RAM OREN and GABRIEL G. KATUL. (2003) Reduction of forest floor respiration by fertilization on both carbon dioxide-enriched and reference 17-year-old loblolly pine stands. Global Change Biology 9:6, 849-861 • Christian P. Andersen. (2003) Sourcesink balance and carbon allocation below ground in plants exposed to ozone. New Phytologist 157:2, 213-228 • Markus reichstein John D. Tenhunen Olivier Roupsard Jean-marc ourcival Serge Rambal Franco miglietta Alessandro peressotti Marco pecchiari Giampiero tirone and Riccardo valentini . (2002) Severe drought effects on ecosystem CO2 and H2O fluxes at three Mediterranean evergreen sites: revision of current hypotheses?. Global Change Biology 8:10, 999-1017 • Monique Carnol Laure Hogenboom M. Ewa Jach Jean Remacle and Reinhart Ceulemans . (2002) Elevated atmospheric CO2 in open top chambers increases net nitrification and potential denitrification. Global Change Biology 8:6, 590-598 • Kurt S. Pregitzer. (2002) Fine roots of trees a new perspective. New Phytologist 154:2, 267-270 • Joseph M. Craine, David A. Wedin and Peter B. Reich . (2001) The response of soil CO2 flux to changes in atmospheric CO2, nitrogen supply and plant diversity. Global Change Biology 7:8, 947-953 • G. A. Bauer, G. M. Berntson and F. A. Bazzaz. (2001) Regenerating temperate forests under elevated CO2 and nitrogen deposition: comparing biochemical and stomatal limitation of photosynthesis. New Phytologist 152:2, 249266 • P.C.D. Newton, H. Clark, G.R. Edwards & D.J. Ross. (2001) Experimental confirmation of ecosystem model predictions comparing transient and equilibrium plant responses to elevated atmospheric CO2. Ecology Letters 4:4, 344-347 Evidence that decomposition rates of organic carbon in mineral soil do not vary with temperature CHRISTIAN P. GIARDINA* AND MICHAEL G. RYAN† Nature 404, 858 - 861 (20 April 2000); doi:10.1038/35009076 Department of Natural Resources and Environmental Management, University of Hawaii at Manoa, 1910 East-West Road, Honolulu, Hawaii 96822 , USA † United States Department of Agriculture-Forest Service, Rocky Mountain Research Station, 240 West Prospect Street, Fort Collins, Colorado 80526, USA, and Graduate Degree Program in Ecology, Colorado State University, Fort Collins, Colorado 80523, USA Correspondence should be addressed to C.P.G. (e-mail: giardina@hawaii.edu). It has been suggested that increases in temperature can accelerate the decomposition of organic carbon contained in forest mineral soil (Cs ), and, therefore, that global warming should increase the release of soil organic carbon to the atmosphere1-6. These predictions assume, however, that decay constants can be accurately derived from short-term laboratory incubations of soil or that in situ incubations of fresh litter accurately represent the temperature sensitivity of C s decomposition. But our limited understanding of the biophysical factors that control Cs decomposition rates, and observations of only minor increases in Cs decomposition rate with temperature in longer-term forest soil heating experiments7-12 and in latitudinal comparisons of Cs decomposition rates13-15 bring these predictions into question. Here we have compiled Cs decomposition data from 82 sites on five continents. We found that Cs decomposition rates were remarkably constant across a global-scale gradient in mean annual temperature. These data suggest that Cs decomposition rates for forest soils are not controlled by temperature limitations to microbial activity, and that increased temperature alone will not stimulate the decomposition of forest-derived carbon in mineral soil. To examine the long-term influence of temperature on the decomposition of Cs in forest soils, we assembled results from studies that used one of two standard methods for estimating Cs loss from soil. Method 1 studies estimate Cs loss by measuring in situ changes in the 13C/12C ratio and total Cs content of soil after existing vegetation is replaced with vegetation that uses a different photosynthetic pathway (for example, C3 forest to C4 pasture)16. The change in vegetation alters the 13C/12C ratio of new detritus, allowing an estimate of the loss of Cs formed before conversion. Method 2 studies estimate Cs loss by incubating soils in the laboratory for 1 yr at temperatures representative of field conditions, and quantifying the CO2 evolved17, 18. The 82 sites examined here span 8 soil orders and the global range of mean annual temperature (MAT) for forests (see Supplementary Information). Calculated Cs turnover times for method 1 studies were unrelated to MAT (R2 = 0.01, P = 0.50; Fig. 1); turnover times under moist tropical conditions were similar to those in cool temperate soils. In these studies, soil clay content, which is thought to control Cs storage1, did not explain the lack of a relationship between Cs decomposition rates and MAT. For soils with similar (15– 27%) clay content, Cs turnover was still unrelated to MAT (R2 = 0.05, P = 0.28). In method 2 studies, Cs turnover time was positively related to incubation temperature, with Cs decomposing more slowly at higher temperatures (R2 = 0.14, P = 0.02; Fig. 2). The Cs lost from method 1 and method 2 soils is the most active, and therefore the most temperature-sensitive, carbon in mineral soil2, 9, 19. However, the decomposition rates of forest-derived Cs reviewed here are insensitive to temperature, unlike the response predicted by models1-6. Figure 1 Relationship between turnover time for mineral soil carbon and mean annual temperature for method 1 studies. Full legend High resolution image and legend (11k) Figure 2 Relationship between observed or modelled Cs turnover time and incubation temperature for incubated forest soils. Full legend High resolution image and legend (14k) In our calculations of turnover time, we assumed a single-pool model for soil carbon. To test whether this assumption affected the results, we compared Cs mass loss per year across MAT, with method 1 sites segregated by time since conversion. For method 1 studies, C s mass loss per year decreased with increasing MAT for sites sampled <11 yr after conversion (R2 = 0.66; P < 0.01), and was unrelated to MAT for sites sampled 11–45 yr after conversion (R2 = 0.01; P = 0.72) or >45 yr after conversion (R2 = 0.15; P = 0.35). Despite a 20 °C gradient in MAT, Cs mass loss as a function of time since conversion was insensitive to temperature (Fig. 3). Cs mass loss was roughly constant with time until about 30 years or about 60% mass loss. After this time, Cs decomposition appears to slow dramatically or stop. For method 2 studies, Cs mass loss for the 1yr incubations decreased with incubation temperature (R2 = 0.19, P < 0.01). We conclude that our choice of a single-pool model did not cause the lack of a relationship between Cs decomposition rates and temperature. Figure 3 Relationship between Cs mass loss and time (in years) since conversion of vegetation type. Full legend High resolution image and legend (14k) Our results conflict with those of Trumbore et al.2, who used 14C-based estimates of light-fraction Cs age (separated from generally smaller quantities of heavy-fraction C s during density fractionation of total Cs) to model the decomposition of light-fraction Cs across a gradient in MAT. These authors found that decomposition rates of light-fraction Cs increased exponentially with increasing temperature. An explanation for the discrepancy with our findings may be that in the study of Trumbore et al., other factors that alter Cs decomposition (moisture, disturbance, and litter quality20, 21) were highest at sites with the highest decomposition rates. The warmer sites were generally wetter, more disturbed, and supported vegetation that produced higher-quality litter. Although many short-term studies of Cs or litter decomposition show that decomposition rates increase with temperature3, 6, 10, transient responses to increasing global temperature are unlikely to represent the response of most detrital carbon in forests. First, detrital carbon in forests resides primarily in the mineral soil (up to 70% in boreal forests13 and 95% in the lowland tropics22, 23), and in situ Cs mass loss rates are much slower than losses of fresh litter or forest-floor material13, 23 . Second, all method 2 studies show large, rapid declines in decomposition rates in the first weeks of incubation, during which <5% of total Cs is typically released9-12, 18, 21. These declines indicate either the depletion of a very small, active pool of Cs, or—because soils are processed before incubation—a return to pre-disturbance conditions. Third, long-term incubations of forest soil9-12, and in situ comparisons of Cs content in heated and unheated soil8 or Cs turnover along gradients in MAT13-15, show responses to increased temperature that are small, ephemeral or nonexistent (that is, Q10 values of 1.0–1.4, where Q 10 = reaction rate at T + 10 °C/reaction rate at T, and T is temperature). Taken together, these data and the data presented in Figs 1 and 2 suggest that sustained, temperature-related increases in the decomposition rate of forest-derived Cs should not be expected. A global-scale relationship between Cs decomposition rates and MAT is central to predictions that global warming will accelerate the release of carbon stored in mineral soil1-6. However, decomposition is performed by enzymes, and enzyme activity is limited by temperature only when the supply rate of substrate exceeds the reaction rate for that substrate. Therefore, the most tenable explanation for the apparent temperature insensitivity of Cs decomposition is that heterotrophic microbes in mineral soil (those organisms responsible for decomposing C s) survive on a supply of substrate that is sub-optimal for growth7, 21, 24. Soil clay content, available moisture, and Cs quality are three factors that may influence substrate availability1, 21. The binding of Cs to clay particles and physical protection within soil aggregates are thought to lower Cs availability1, 21. At a given location, where variations in climate and biota are more uniform, turnover times are longer for Cs associated with clay than with sand20, 25. If clay controls substrate availability, we would expect Cs turnover times to increase as clay content increases. However, data from method 1 and method 2 studies provide only weak support for clay limitations to substrate availability. For both method 1 and method 2 studies, Cs turnover time was nearly constant across sites, and variation among soils within a region was similar to global variation. For method 1 studies, where clay varied from 7% to 84%, there was no relationship between Cs turnover time and soil clay content (R 2 = 0.05, P = 0.14; see Supplementary Information ). In method 2 studies, Cs mass loss was weakly related to clay content in soils with 7–30% clay (R2 = 0.15, P = 0.05; Supplementary Information). For clay >30%, Cs mass loss decreased with increased clay (R2 = 0.41, P = 0.06; Supplementary Information). While these results conflict with established modelling assumptions1, other studies have also found weak relationships between clay content and Cs decomposition rates17, 26. Available moisture exerts a large influence on soil microbial activity21, and low soil moisture probably reduces microbial populations. However, available moisture did not affect substrate availability in method 2 studies because soils were maintained at or near field moisture capacity for the length of the incubation. Although we have no information on differences in soil moisture among method 1 sites, method 1 data were taken primarily from sites that had previously supported closed-canopy forests, which indicates moisture regimes that are at least adequate for decomposition. Low Cs quality may limit substrate availability for microbes, and perhaps also limit microbial populations21. Forest-derived Cs consists of lignin-dominated remains and precipitated byproducts of plant and microbial residue decomposition. These compounds are poor carbon sources for microbes, because energy yields are low for the energy expended to digest them 21. Evidence that low quality of Cs limits Cs decomposition rates in mineral soil includes low C s decomposition rates compared with rates for fresh litter23 and rapid increases in CO2 release from nonrhizosphere soils that are amended with labile substrate24. 13C nuclear magnetic resonance analyses of forest-derived Cs show that the relative abundance of Cs functional groups (for example, alkyls, O-alkyls, aromatics and carbonyls) varies minimally across global-scale gradients in MAT (ref. 27). If Cs quality limits decomposition rates, then low global-scale variation in Cs decomposition rates may reflect low variation in the chemical composition of Cs. Whether Cs decomposition is controlled by temperature or by substrate availability will change predictions for the effect of global warming on the large quantity of Cs stored in tropical soils1, 2, 28 and in soils that are frozen for most of the year28, 29. If temperature limits Cs decomposition, as assumed in most current ecosystem models, then tropical soils would provide the main source of additional carbon released in a warmer climate because Cs decomposition rates in high-latitude soils would be constrained by perennially low soil temperatures1, 2. In contrast, if substrate availability limits Cs decomposition rates, increased global temperatures alone would have little influence on Cs decomposition rates in the tropics. Warming at high latitudes, however, would expose larger amounts of Cs to microbial activity by lowering the depth of frozen soil, lowering the water table, and extending the duration of thawed conditions29. Once these soils are thawed or the water table lowered, Cs decomposition rates would be constrained by substrate quality and moisture rather than by low temperatures. The turnover of Cs in forest soils appears to be remarkably constant on a global scale, and insensitive to differences in MAT. However, the relationship between Cs turnover and MAT presented here serves only as a proxy for the changes that will occur in situ in response to global warming. The influence of temperature on Cs decomposition rates needs to be directly examined across a range of sites to better constrain predictions of the effects of warming on carbon release from soil. Methods Method 1 studies primarily examined the loss of forest-derived Cs . All sites were disturbed during conversion, but land management varied during and following conversion. Details on methodology for a typical method 1 study are given in ref. 16. For method 2 studies, all soils were sampled from closed-canopy forests, similarly processed, and maintained at constant moisture levels near field capacity for the length of the incubation period. We used studies in which soils were incubated for at least 1 yr, because long-term incubations provide information on the mineralization potential of both the small labile Cs pool and the larger intermediate Cs pool9. Microbial activity and Cs decomposition rates also may be less sensitive to sampling disturbance in long-term incubations than in short-term incubations. Method 2 studies have some important limitations. Soil macro-structure is altered during sampling and processing. Soils are incubated at a constant temperature and high moisture, whereas field environments fluctuate and are generally drier. Finally, incubations eliminate carbon inputs, which can shift microbial populations 18. Although incubation-derived indices of Cs decomposition rates will probably differ from those developed in the field, relative differences among sites are presumably real17, 18, 21, 26. We used single-pool exponential decay models to estimate Cs turnover time in method 1 and method 2 studies. Single-pool models are widely used to describe Cs turnover, but they may overestimate rates because the approach assumes that all Cs will behave as did the Cs released during the study30. It is unlikely, however, that our single-pool approach masked a relationship between the turnover of a large, temperature-sensitive Cs pool and temperature. In method 1 studies, we would suspect masking if Cs mass loss per year increased with temperature early in the decomposition sequence. However, for sites with <11 yr since conversion, Cs mass loss per year was higher for cool sites (9.5% yr-1) than for warm (4.5% yr-1 ) climate soils (P < 0.01). In method 2 studies, we would suspect masking if the total quantity of Cs released per gram of soil increased with temperature because the release of this C s is independent of total Cs pool size. However, Cs released per kilogram of soil in one year declined with increasing incubation temperature (R2 = 0.11, P = 0.04). Further, for masking to have occurred, the proportion of total Cs that is temperature sensitive or fast-cycling must decline steeply with increasing temperature. Using the best techniques available, Trumbore et al.2 found no relationship between the proportion of fastcycling C s (from 50% to 80% of total Cs) and MAT. Nonetheless, we tested the potential for masking by assuming that all Cs released in method 2 studies was fast cycling and that 50%, 65% and 80% of total C s in tropical, temperate and subalpine soils, respectively, was fast cycling. We then recalculated turnover times for fast-cycling Cs alone. Overall, patterns of Cs decomposition were unchanged. Supplementary information accompanies this paper. Received 10 August 1999; accepted 14 March 2000 References 1. Schimel,D. S. et al. Climatic, edaphic, and biotic controls over storage and turnover of carbon in soils. Glob. Biogeochem. Cycles 8, 279-293 (1994). | Article | ISI | ChemPort | 2. Trumbore,S. E., Chadwick,O. & Amundson,A. Rapid exchange between soil carbon and atmospheric carbon dioxide driven by temperature change. Science 272, 393-396 (1996). | ISI | ChemPort | 3. Kirschbaum,M. U. F. 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 (1995). | Article | ISI | ChemPort | 4. VEMAP members. Vegetation/ecosystem modeling and analysis project: Comparing biogeography and biogeochemistry models in a continental-scale study of terrestrial ecosystem response to climate change and CO2 doubling. Glob. Biogeochem. Cycles 9, 407-437 (1995). | ISI | 5. Cao,M. & Woodward,F. I. Dynamic responses of terrestrial ecosystem carbon cycling to global climate change. Nature 393, 249-252 (1998). | Article | ISI | ChemPort | 6. Jenkinson,D. S., Adams,D. E. & Wild,A. Model estimates of CO2 emissions from soil in response to global warming. Nature 351, 304-306 (1991). | Article | ISI | ChemPort | 7. Nadelhoffer,K. J., Giblin,A. E., Shaver,G. R. & Laundre,J. A. Effects of temperature and substrate quality on element mineralization in six arctic soils. Ecology 72, 242-253 (1991). | ISI | 8. Peterjohn,W. T., Melillo,J. M., Bowles,F. P. & Steudler,P. A. Soil warming and trace gas fluxes: experimental design and preliminary flux results. Oecologia 93, 18-24 (1993). | ISI | 9. Townsend,A. R., Vitousek,P. M., Desmarais,D. J. & Tharpe,A. Soil carbon pool structure and temperature sensitivity inferred using CO2 and 13CO2 incubation fluxes from five Hawaiian soils. Biogeochemistry 38, 1-17 (1997). | Article | ISI | ChemPort | 10. Winkler,J. P., Cherry,R. S. & Schlesinger,W. H. The Q10 relationship of microbial respiration in a temperate forest soil. Soil Biol. Biochem. 28, 1067-1072 (1996). | Article | ISI | ChemPort | 11. Updegraff,K., Pastor,J., Bridgham,S. D. & Johnston,C. A. Environmental and substrate controls over carbon and nitrogen mineralization in northern wetlands. Ecol. Applic. 5, 151-163 (1995). | ISI | 12. Holland,E. A., Townsend,A. R. & Vitousek,P. M. Variability in temperature regulation of CO 2 fluxes and N mineralization from five Hawaiian soils: implications for a changing climate. Glob. Change Biol. 1, 115-123 (1995). | ISI | 13. Liski,J., Livesniemi,H., Maketa,A. & Westman,C. CO2 emissions from soil in response to climate warming are over-estimated—The decomposition of old soil organic matter is tolerant of temperature. Ambio 28, 171-174 (1999). | ISI | 14. Feller,C. & Beare,M. H. Physical control of soil organic matter dynamics in the tropics. Geoderma 79, 69-116 (1997). | Article | ISI | ChemPort | 15. Gregorich,E. G., Ellert,B. H. & Monreal,C. M. Turnover of soil organic matter and storage of corn residue C estimated from C-13 abundance. Can. J. Soil Sci. 75, 161-167 (1995). | ISI | ChemPort | 16. Bashkin,M. & Binkley,D. Changes in soil carbon following afforestation in Hawaii. Ecology 79, 828-833 (1998). | ISI | 17. Motavalli,P. P. et al. Comparison of laboratory and modeling simulation methods for estimating soil carbon pools in tropical forest soils. Soil Biol. Biochem. 26, 935-944 (1994). | Article | ISI | 18. Hart,S. C., Nason,G. E., Myrold,D. D. & Perry,D. A. Dynamics of gross nitrogen transformations in an old-growth forest: the carbon connection. Ecology 75, 880-891 (1994). | ISI | 19. Trumbore,S. E. et al. Belowground cycling of carbon in forests and pastures of Eastern Amazonia. Glob. Biogeochem. Cycles 9, 515-528 (1995). | Article | ISI | ChemPort | 20. Balesdent,J., Mariotti,A. & Guillet,B. Natural 13C abundance as a tracer for studies of soil organic matter dynamics. Soil Biol. Biochem. 19, 25-30 (1987). | Article | ISI | ChemPort | 21. Paul,E. & Clark,F. Soil Microbiology and Biochemistry (Academic, New York, 1996). 22. Binkley,D. & Resh,S. C. Rapid changes in soils following Eucalyptus afforestation in Hawaii. Soil Sci. Soc. Am. J. 63, 222-225 (1999). | ISI | ChemPort | 23. Veldkamp,E. Organic carbon turnover in three tropical soils under pasture after deforestation. Soil Sci. Soc. Am. J. 58, 175-180 (1994). | ISI | 24. Cheng,W. et al. Is available carbon limiting microbial respiration in the rhizosphere? Soil Biol. Biochem. 28, 1283-1288 (1996). | Article | ISI | ChemPort | 25. Bonde,T., Christensen,B. T. & Cerri,C. C. Dynamics of organic matter as reflected by natural 13C abundance in the particle size fractions of forested and cultivated Oxisols. Soil Biol. Biochem. 24, 275277 (1992). | Article | ISI | ChemPort | 26. Sørenson,L. H. Carbon-nitrogen relationships during the humification of cellulose in soils containing different amounts of clay. Soil Biol. Biochem. 13, 313-321 (1981). | Article | ISI | 27. Mahieu,N., Powlson,D. H. & Randall,W. Statistical analysis of published carbon-13 CPMAS spectra of soil organic matter. Soil Sci. Soc. Am. J. 63, 307-319 (1999). | ISI | ChemPort | 28. Post,W. M., Emanuel,W. R., Zinke,P. J. & Stangenberger,A. G. Soil carbon pools and world life zones. Nature 298, 156-159 (1982). | ISI | ChemPort | 29. Goulden,M. L. et al. Sensitivity of boreal forest carbon balance to soil thaw. Science 279, 214-217 (1998). | Article | PubMed | ISI | ChemPort | 30. Townsend,A. R., Vitousek,P. M. & Trumbore,S. E. Soil organic matter dynamics along gradients in temperature and land use on the Island of Hawaii. Ecology 76, 721-733 (1995). | ISI | Acknowledgements. We thank R. Hubbard, D. Binkley, R. Waring, S. Hart, S. Trumbore, I. Døckersmith, R. Sanford Jr and M. Bashkin for comments on earlier versions of this manuscript, and R. King for help with statistics. This work was supported by the US National Science Foundation and the USDA Forest Service. © Privacy Policy 2000 Nature Publishing Group Plant First Physiology published on December 30, Preview 2003; 10.1104/pp.103.030569 Plant Physiology, January 2004, Vol. 134, pp. 520-527 ENVIRONMENTAL STRESS AND ADAPTATION Respiratory Oxygen Uptake Is Not Decreased by an Instantaneous Elevation of [CO2], But Is Increased with Long-Term Growth in the Field at Elevated [CO2]1 Phillip A. Davey2, Stephen Hunt, Graham J. Hymus3, Evan H. DeLucia, Bert G. Drake, David F. Karnosky and Stephen P. Long* Departments of Crop Sciences and Plant Biology, University of Illinois, Urbana, Illinois 61801 (P.A.D., S.P.L., E.H.D.); Department of Biology, Queen’s University, Kingston, Ontario, K7L 3N6, Canada (S.H.); Smithsonian Environmental Research Center, Edgewater, Maryland 21307 (G.J.H., B.R.D.); and School of Forest Resources and Environmental Science, Michigan Technological University, Houghton, Michigan 49931 (D.F.K.) ABSTRACT Averaged across many previous investigations, doubling the CO2 concentration ([CO2]) has frequently been reported to cause an instantaneous reduction of leaf dark respiration measured as CO2 efflux. No known mechanism accounts for this effect, and four recent studies have shown that the measurement of respiratory CO2 efflux is prone to experimental artifacts that could account for the reported response. Here, these artifacts are avoided by use of a high-resolution dual channel oxygen analyzer within an open gas exchange system to measure respiratory O2 uptake in normal air. Leaf O2 uptake was determined in response to instantaneous elevation of [CO2] in nine contrasting species and to long-term elevation in seven species from four field experiments. Over six hundred separate measurements of respiration failed to reveal any decrease in respiratory O2 uptake with an instantaneous increase in [CO2]. Respiration was found insensitive not only to doubling [CO2], but also to a 5-fold increase and to decrease to zero. Using a wide range of species and conditions, we confirm earlier reports that inhibition of respiration by instantaneous elevation of [CO2] is likely an experimental artifact. Instead of the expected decrease in respiration per unit leaf area in response to long-term growth in the field at elevated [CO2], there was a significant increase of 11% and 7% on an area and mass basis, respectively, averaged across all experiments. The findings suggest that leaf dark respiration will increase not decrease as atmospheric [CO2] rises. A quantitative analysis of prior studies weighted for replication and experimental variation concluded that a doubling of atmospheric CO2 concentration would decrease respiratory CO2 efflux by 18% in woody plants (Curtis and Wang, 1998 ). This decrease is considered the result of a direct instantaneous effect of increased CO2 concentration and a longer term indirect effect due to changes in leaf composition with long-term growth at elevated CO2 (for review, see Drake et al., 1999 ). In an analysis of 45 species, 36 showed an average 15% instantaneous reduction in net respiratory CO2 efflux per unit leaf area (Rd,CO2) on transfer to elevated [CO2] (for review, see Amthor, 1997 ). Terrestrial plant respiration releases 10 times more carbon per annum than fossil fuel combustion (Amthor, 1997 ). Therefore, a 15% to 20% decrease in foliar respiration might increase the potential of terrestrial vegetation to sequester carbon into biomass, providing a partial amelioration to the rate of increase of atmospheric [CO2] (for review, see Gonzalez-Meler and Siedow, 1999 ). However, there is considerable variation in the reported instantaneous effects of elevated [CO2] on Rd,CO2 with some studies reporting a large decrease and others reporting no change (Bunce, 2002 ; Bruhn et al., 2002 ). Furthermore, whereas some biochemical reactions within plant respiratory metabolic pathways are sensitive to [CO2], none of those identified exert sufficient metabolic control to account for the reported decreases in respiration (Gonzalez-Meler and Siedow, 1999 ). Four developments now suggest that the decreases observed when measuring Rd,CO2 from leaves in gas exchange systems incorporating infrared CO2 analyzers may be the result of artifacts in the measuring system. Amthor (2000 ) showed that very thorough sealing and use of an enlarged cuvette enclosing the leaf greatly reduced the apparent instantaneous effect of doubling [CO2] on Rd,CO2, although a small significant effect still remained. Jahnke (2001 ) systematically analyzed and eliminated potential artifacts within a gas exchange system and subsequently failed to detect any response in bean (Phaesoleus vulgaris) and poplar (Populus tremula) leaves when corrections were applied. Bouma et al. (1997 ) and Burton and Pregitzer (2002 ) similarly demonstrated that respiratory efflux of CO2 from roots was independent of measurement CO2 concentration. Assuming common dark respiratory metabolism between roots and leaves, this casts further doubt on the possibility of a direct response in the leaf. Avoiding the potential artifacts of CO2 exchange measurement by determining dark respiratory O2 uptake (Rd,O2), Amthor et al. (2001 ) showed no significant change in respiration of five Rumex crispus leaves in sharp contrast to earlier measurements of Rd,CO2 in the same species (Amthor et al., 1992 ). Measurement of O2 uptake has three important advantages over measurement of CO2 efflux for an instantaneous effect of change in [CO2] on respiration. First, the gas being measured is not the gas being altered in concentration, avoiding any need for instrument recalibration. Second, the concentration gradient of [O2] between the cuvette enclosing the leaf and the surrounding air is unaltered when [CO2] is changed. Finally, O2, unlike CO2, is not easily absorbed and adsorbed by surfaces in the gas exchange system. Yet, even when the effect of elevated [CO2] on O2 uptake has been examined, findings have been variable. In contrast to Amthor et al. (2001 ), Reuveni et al. (1993 ) measured a decrease in respiratory O2 uptake under elevated [CO2]. Controversy over the presence or absence of a direct effect of [CO2] on Rd,CO2 continues. Some recent studies of Rd,CO2, despite controlling for errors likely to arise from leaks, continue to report a significant instantaneous decrease in Rd,CO2 on elevation of [CO2]. For example, decreases have been reported of 19% for rice (Oryza sativa) canopies on elevation of [CO2] from 350 to 700 µmol mol-1 (Baker et al., 2000 ); 31% to 79% for Scots pine (Pinus sylvestris) needles also on elevation from 370 to 700 µmol mol-1 (Jach and Ceulemans, 2000 ); 16% to 40% for eight species of crop and crop weeds on increase from 60 to 1,000 µmol mol-1 (Bunce, 2001 ); an 11% average for leaves of 12 grass species on elevation from 360 to 1,300 µmol mol-1 (Tjoelker et al., 2001 ); 18% in the second leaf of soybean (Glycine max) on elevation from 350 to 1,400 µmol mol-1 (Bunce, 2002 ); and 4.5% decrease for seedling shoots of the tropical rainforest tree Tectona grandis on elevation from 370 to 600 µmol mol-1 (Holtum and Winter, 2003 ). Precautions taken in measurement were sufficient for Bunce (2002 ) to rebut the idea that the direct effect of elevated [CO2] on respiration was an artifact: “These precautions coupled with observed effects of dark [CO2] treatments on mass accumulation and translocation, suggest that the observed changes in respiration were real.” Although these decreases are smaller than those suggested by earlier studies, and some use larger increases in [CO2] to elicit an effect, they continue to suggest direct instantaneous inhibition of leaf dark respiration by increase in [CO2]. An added uncertainty is whether an instantaneous suppression of respiration by elevated [CO2] is limited to certain species or conditions (Drake et al., 1999 ). Although maintaining the same [CO2] outside the cuvette to that inside will minimize diffusive leaks, it will not eliminate the substantial problem of CO2 absorption/desorption from surfaces in the gas exchange system (Bloom et al., 1980 ; Long et al., 1996 ; Jahnke, 2001 ) or the mass flow of CO2 between the cuvette and the outside through the intercellular air space when only part of the leaf is enclosed (Jahnke and Krewitt, 2002 ). Thus, some of these further artifacts may still affect even these very careful studies. Clearly, we are some way from any consensus over the existence, or otherwise, of a direct effect of increase in [CO2] on leaf respiration in the dark. If the artifacts suggested by Amthor (2000 ), Jahnke (2001 ), and Jahnke and Krewitt (2002 ) apply to the effect of an instantaneous increase in [CO2] on Rd,CO2, then the same errors would at least in part apply when these systems are used to measure the effect of long-term acclimation to elevated [CO2] on Rd,CO2. Plants grown in elevated [CO2] in the long-term (1-10 years) typically show decreased nitrogen content and increased photosynthesis, growth, leaf mass per unit area, nonstructural carbohydrate concentrations (Drake et al., 1997 ), and large increases in leaf mitochondrial numbers (Robertson et al., 1995 ; Griffin et al., 2001 ; Tissue et al., 2002 ). With the exception of nitrogen content, these are all factors that might be expected to lead to increased leaf respiration. However, a quantitative survey using meta-analytical analysis concluded that long-term growth at elevated [CO2] decreased Rd,CO2 significantly and by 18% (Wang and Curtis, 2002 ). Recently, high-sensitivity dual lead-gold (fuel cell) detectors have been developed that allow measurement of small differences in [O2] (approximately 1-2 µmol mol-1) between the inlet and outlet air streams of a leaf cuvette in normal air (Willms et al., 1997 , 1999 ; Hunt, 2003 ). The objectives of this study were to use such a measurement system to answer two questions: Can the findings of Amthor et al. (2001 ) of no instantaneous effect of doubling of [CO2] on leaf Rd,O2 be extended to a wide range of species, growth conditions, and measurement [CO2]? Using a range of long-term experiments in which a wide range of plants have been grown under elevated [CO2] in the field (Table I) is Rd,O2 decreased or unaffected as a result of acclimation to elevated [CO2]? View this table: [in this window] [in a new window] Table I. Species, [CO2] treatments, plant and leaf developmental stage at the time of measurement Locations for material growing in the field and microclimate for controlled environment grown plants are given under “Materials and Methods”. References providing a description of each of the long-term field elevated [CO2] experiments used are included. RESULTS There was no significant instantaneous effect of [CO2] on leaf Rd,O2 in any of the nine species studied (Fig. 1; Tables II) and III). The results for the ANOVA were, for CO2, F3,332 = 0.032, P = 0.99; and for species x CO2, F24,332 = 0.528, P = 0.97. A complete lack of any effect of [CO2] was evident in all species not only on doubling [CO2], but when [CO2] was increased more than five times above the current ambient concentration or decreased to zero (Fig. 1; Tables II and III). The extremely high probabilities for accepting the null hypothesis (P > 0.97) eliminate any possibility of a Type II error, i.e. a difference undetected because of variability. This is also reflected in the combined regression of response relative to ambient [CO2] (Fig. 1), where the probability that the response is independent of [CO2] was P = 0.99 (F3,360 < 0.001). This lack of response of respiration to elevated [CO2] was independent of treatment method (F3,360 = 0.222, P = 0.88), developmental stage (F3,360 = 0.174, P = 0.92), beginning or end of night (F3,360 = 0.638, P = 0.59), and the [CO2] at which the plants had been grown (F3,100 = 0.080, P = 0.97). Although there was no interaction with [CO2], absolute rates of respiration generally decreased with time in the dark. Figure 1. The instantaneous effect of change in [CO2], from the current ambient concentration, on leaf respiratory O2 uptake in the dark. Respiration is plotted as the percentage of change (±1 SE) relative to the rate of O2 consumption at current ambient [CO2]. Each mean illustrated is based on measurements of 274 plants and nine species (Table I). Leaves were in darkness and their temperature was maintained at 25°C. Results for an analysis View larger version (13K): of variance (ANOVA) are given in the results text, and [in this window] respiration rates for individual species contributing to the [in a new window] illustrated means are shown in Tables II and III. View this table: [in this window] [in a new window] View this table: [in this window] [in a new window] Table II. The response of leaf respiration to changes in [CO2] for plants grown at current [CO2] only. The mean dark leaf O2 uptake on an area basis (micromoles per minute per second), and its SE (n = 6), are given for each of the four species at the beginning and end of the 10-h night. Table III. The response of leaf respiration to changes in (CO2) for species grown at long-term elevated (CO2) experiments The mean respiration rate and its SEM are given on an area basis and a dry weight basis. n indicates the number of plots or open-top chambers from which measurements of each treatment were made. Bold values show the critical longterm comparison of leaves grown and measured at their growth concentrations. An asterisk indicates that leaves grown and measured at elevated (CO2) have significantly different rates (P < 0.05) to those grown and measured at current ambient (CO2). Despite the lack of an instantaneous effect, long-term growth at elevated [CO2] did affect respiration. Leaves grown and measured at elevated [CO2] had the same or a significantly higher Rd,O2 than those grown and measured at current ambient [CO2] (Table III). The largest increases in respiration were for Quercus geminata (23%) and soybean (22%) on a leaf area basis. By contrast, there was no increase in Acer saccharum, Betula papyrifera, and Populus tremuloides (Table IIIa). When all values from the long-term experiments were combined, there was a significant increase in Rd,O2 of 11% and 7% on a leaf area and mass basis, respectively, and relative to controls (Fig. 2). Figure 2. The effect of long-term (lifetime or >3 years) growth at elevated [CO2] in the field on leaf respiration. The data points show mean leaf respiratory O2 uptake (±90% confidence intervals) on an area and mass basis for View larger version (8K): leaves grown at elevated [CO2] relative to controls grown at [in this window] current ambient [CO2]. The points are means for seven [in a new window] species (Table I) weighted for replication and variance using meta-analytical statistical methods. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED DISCUSSION Six hundred separate measurements of different leaves, encompassing four long-term field CO2 enrichment experiments revealed no evidence of any decrease in Rd,O2 in response to instantaneous elevation of [CO2]. This was consistent for seven tree and two crop species, the majority sampled from long-term elevated [CO2] experiments where plants have developed throughout their life or for several seasons under elevated [CO2]. This lack of response was irrespective of developmental stage, growth conditions, and growth [CO2]. Bunce (2001 ) reported that the sensitivity of Rd,CO2 to an instantaneous elevation of [CO2] increased with duration of the dark period. However, no effect on Rd,O2 was evident here (Tables II and III), regardless of whether measurements were made at the start or end of the natural dark period. Jahnke (2001 ) similarly found no effect of [CO2] on Rd,CO2 even after 72 h of darkness. Tjoelker et al. (2001 ) noted that the instantaneous decrease in Rd,CO2 with elevated [CO2], although minor (1.8%) in response to increase from 360 to 700 µmol mol-1, was far more substantial (11%) on increase from 360 to 1,300 µmol mol-1. However, a 5-fold increase above current ambient [CO2] or removal of CO2 from the atmosphere around the leaf did not alter Rd,O2 (Fig. 1). These findings support the conclusions of Amthor et al. (2001 ) and of Jahnke (2001 ) and extend them to more species and to major long-term [CO2] field manipulation studies. Although the results show no effect of instantaneous elevation of [CO2] on Rd,O2, this cannot rule out some effect on apparent Rd,CO2 due to non-respiratory processes, for example, CO2 uptake into carboxylic acids. However, in one of the very few studies to measure CO2 and O2 exchange at different [CO2], Amthor et al. (2001 ) found no effect of elevated [CO2] on respiratory quotient. It has been suggested that the variability of response to instantaneous increase in [CO2] apparent in the literature may be the result of species differences (Drake et al., 1999 ; Baker et al., 2000 ; Hamilton et al., 2001 ). Without definitive testing, it is impossible to be certain on this point. However, this study contains a diverse range of species and functional types (including C3 and C4 photosynthetic types, herbaceous and woody forms, two major crops, and seven forest dominant gymnosperm and angiosperm species) that individual exceptions exist now seem somewhat improbable. Moreover, we deliberately selected species previously reported to show some of the largest inhibitions of respiration by elevated CO2 with previous methods and technology: soybean 40% (Bunce, 1995 ), Pinus taeda 14% (Teskey, 1995 ), Quercus rubra 5.6% (Amthor, 2000 ), and maize (Zea mays) 46% (Cornic and Jarvis, 1972 ). Any errors associated with measuring the instantaneous response of respiratory CO2 efflux to elevated [CO2] may be equally applicable to measurement of the long-term effects of growth at elevated [CO2] on Rd,CO2. The significant 7% increase in Rd,O2 on a mass basis, averaged across seven species from four long-term experiments (Fig. 2), contrasts sharply with the average 18% decrease reported by Wang and Curtis (2002 ) in their meta-analysis of Rd,CO2 on a mass basis measured in long-term experiments. The significant increase on a mass, as well as an area basis, suggests that increased respiration per unit leaf area is not just the result of the increased mass per unit area that is commonly observed in leaves grown at elevated [CO2] (Drake et al., 1999 ). However, as noted above, this mean increase masks marked differences between species in this long-term response (Table III). To our knowledge, this is the first study to use this alternative method to investigate the effects of long-term growth at elevated [CO2] on respiration. The increase in respiration shown here is consistent with the widely reported increase in leaf-soluble carbohydrates with long-term growth at elevated [CO2] (Moore et al., 1999 ). Azcon-Bieto and Osmond (1983 ) showed that increase in leaf nonstructural carbohydrate content produced by manipulation of photosynthesis during the photoperiod increased subsequent dark respiration rates. Therefore, a similar response might be expected when elevated [CO2] increases photosynthesis and leaf nonstructural carbohydrate content during the photoperiod. In conclusion, instantaneous elevation of [CO2] had no effect on leaf O2 uptake in these nine species. This further confirms the suggestion that reports of respiratory inhibition, based on measurements of CO2 uptake, are the result of experimental artifacts and not the result of any sensitivity of plant respiration to the [CO2] at the time of measurement. A 15% to 20% reduction of terrestrial plant respiration with a doubling of atmospheric [CO2] concentration would represent some amelioration, at least temporarily, in the rate of rise in the global atmospheric [CO2] (for review, see Gonzalez-Meler and Siedow, 1999 ). It can no longer be assumed that the direct effect of elevated [CO2] on plant respiration will reduce future ecosystem CO2 efflux. On the basis of the long-term field manipulations of [CO2] studied here, a small increase in respiration per unit leaf area should be expected—this will amplify the effect of any increase in leaf area due to growth at elevated [CO2]. MATERIALS AND METHODS Growth Conditions The species, developmental stage, growth conditions, and experimental site are listed in Table I. Seven out of nine species were grown at current ambient (368 µmol mol-1) and at an elevated [CO2] (550 or 700 µmol mol-1). Two of the species were measured at more than one developmental stage; soybean (Glycine max) at vegetative and podfill stages, and Pinus taeda at juvenile and mature stages. Soybean was grown in controlled environment- and open-field conditions (FACE). Oxygen Analysis System Leaf O2 uptake rate was determined in an open gas exchange system incorporating a dual-cell oxygen analyzer (S-3A/DOX; AEI Technologies, Pittsburgh) described by Willms et al. (1997 ), and incorporating a stainless steel and glass leaf gas exchange cuvette (MPH-1000; Campbell Scientific, Logan, UT). The analyzer incorporates a novel differential oxygen sensor obtained by modifying two electrochemical fuel cells (model KE-25; Figaro USA, Wilmette, IL), placing one in a reference gas stream and the other in an analytical gas stream, and then ensuring that the two gas sensors were exposed to an identical temperature and pressure (Willms et al., 1997 ). The O2 fuel cells contain a Teflon membrane through which O2 can diffuse, and a weak acid electrolyte bathes a lead anode and a gold cathode electrode. In the environment of the sensor, the O2 oxidizes the lead, forming PbO, which dissolves into the solution and generates a small current. The difference in output of the two sensors is directly proportional to the [O2] between the air streams. The selective diffusive properties of Teflon and the specific electrochemistry ensure that the differential sensor responds directly only to O2 in air. However, changes in partial pressures of CO2 and water vapor due to respiration and transpiration will affect that of O2 causing an indirect effect (Willms et al., 1999 ). For this reason, these gases are removed from the air stream before the O2 sensors (Hunt, 2003 ). The analyzer was secured to a vibrationless table top and additional layers of insulation were wound around the O2 measurement cells and instrument casing to further improve resolution. The differential resolution of the modified dual-cell O2 analyzer was 1 to 2 µmol mol-1 against a background [O2] of 210 mmol mol-1. The cuvette was large enough to enclose the entire leaf, leaflet, or cohort of needles; typically, 50 to 120 cm2 of leaf were enclosed. This avoided errors associated with gas flow between the cuvette and surrounding air via the intercellular air space that Jahnke and Krewitt (2002 ) have shown to result from the use of cuvettes that only enclose part of the leaf. The cuvette (MPH-1000; Campbell Scientific) has been described in detail previously (Bingham et al., 1980 ), adapted from the earlier design of Bingham and Coyne (1977 ). Typical boundary layer conductances to water vapor are 2.5 to 4 mol m-2 s-1 (calculated from Bingham and Coyne, 1977 ; Bingham et al., 1980 ). A gas exchange system (LI6400; LI-COR, Lincoln, NE) was used to provide a controlled [CO2] to the cuvette, and air flow rate was measured by a pair of flow meters (F350; AEI Technologies), calibrated as described previously (Long et al., 1996 ). The gas flow across the leaf was maintained at 2.5 cm3 s-1. This flow rate and the enclosed leaf area resulted in a [O2] of 40 to 80 µmol mol-1. The entry air streams to the oxygen analyzer were scrubbed of water and CO2 by magnesium perchlorate and soda lime. Column sizes were calculated to remove all CO2 and to be sufficient to dry the air stream to the equilibrium vapor pressure of magnesium perchlorate (approximately 0.1 Pa). The efficacy of these columns was tested by switching dry air (<0.01 kPa water vapor) with a humidified air stream (2.0 kPa water vapor) at the cuvette inlet; no change in [O2] from 0 µmol mol-1 was recorded. Similar tests showed that the soda lime column was effective in preventing interference from differences in [CO2] between the cuvette and reference air streams. The response time of the gas exchange system was assessed using two cylinders of compressed air found to differ in [O2] by 80 µmol mol-1. Air from one of these cylinders was fed to the empty cuvette inlet and also the reference air stream until a constant [O2] of 0 µmol mol-1 was obtained at the analyzer. Air from the second cylinder was then switched in at the cuvette inlet and the time taken to obtain a constant [O2] of 80 µmol mol-1 at the analyzer was recorded. The change was 98% complete within 10 min and complete within the resolution of the analyzer at 15 min. Therefore, it was assumed that 20 min was adequate to record any response of respiratory O2 efflux to an instantaneous change in [CO2]. Respiration Measurements Respiration was measured in the last 2.5 h of the night in all species and additional measurements were made in the first 2.5 h of the night for three species (Table I). Measurements were made at each location listed in Table I. However, because of its size and stability, the gas exchange system was housed in a field laboratory at each site and leaves were detached shortly before measurement. Petioles were cut under water and their cut ends were kept immersed in water until measurements were complete. The leaves remained turgid throughout. For pot-grown soybean (Table I), parallel measurements on attached and detached leaves were made; no significant differences in respiration or its response to [CO2] were found. Leaves were equilibrated in the dark to 25°C before measurement. For a single replicate measurement, respiration was measured at a [CO2] of 2,000 µmol mol-1, and then again at 700, 550, 360, and 0 µmol mol-1 CO2. To avoid systematic error, each alternate replicate measurement was started at 0 µmol mol-1 CO2 and was then switched to 368, 700, 550, and 2,000 µmol mol-1 CO2. Species from the FACE experiments at Duke Forest, Rhinelander, and the University of Illinois, and the open-top chamber experiment at Merritt Island were measured at 550 µmol mol-1 [CO2] rather than at 700 µmol mol-1 [CO2] because this was the growth concentration (Table I). After each step-wise change in [CO2], a 20min waiting period ensured adequate time for any instantaneous response in Rd,O2. Each set of measurements took 80 min to complete, and leaves were maintained in the dark at a leaf temperature of 25°C for the entire measurement period. After each set of measurements, leaf area was measured by digital imaging and the leaves were then dried to a constant weight at 80°C. Leaf O2 evolution rates were calculated on a leaf area and a dry mass basis by adapting the CO2 uptake equations of von Caemmerer and Farquhar (1981 ) for open system gas exchange measurements. Statistical Analysis In the four long-term field experiments, replicate number was determined by the number of treatment plots (Table III). This avoided pseudoreplication. A minimum of three repeat measurements were taken in each treatment plot of each experiment, and their pooled values provided the individual sample measure. Only these data were used to assess the effect of longterm growth at elevated [CO2] on respiratory O2 uptake. Outside of these long-term elevated [CO2] studies, measurements were made from six randomly selected plants each of P. taeda seedlings, Q. rubra mature trees, and maize flowering plants growing under current ambient [CO2] in the open. Soybean was grown in two separate controlled environment studies, and here, replicates represent the number of separate chambers used (Table I). ANOVA was used to test the effect of immediate changes of [CO2] on leaf respiration rate for all nine species combined (SYSTAT, Evanston, IL). This tested the effect of species, growth conditions, developmental stage, time of measurement, and growth [CO2] on the instantaneous response to elevated [CO2]. In addition, a regression analysis of response relative to ambient [CO2] was tested (SYSTAT). Because of the very different experimental designs, durations, and treatment procedures of the experiments in which plants were grown at elevated [CO2], and the possibility that species may respond differently to long-term growth in elevated [CO2], the pooled effect of all field studies was assessed by meta-analytical statistical techniques (Hedges et al., 1999 ; Ainsworth et al., 2002 ; Morgan et al., 2003 ) using a software package (MetaWin, Rosenberg et al., 2000). This allows determination of the consensus of studies differing in design. The natural log of the response ratio (respiratory O2 uptake for leaves grown at elevated [CO2]/respiratory O2 uptake for leaves grown at the current ambient [CO2]) was used as the metric for analysis (Hedges et al., 1999 ). A mixed-model analysis was used based on the assumption of random variation in effect sizes between individual measurements. A weighted parametric analysis was used (Gurevitch and Hedges, 1999 ), and each individual observation was weighted by the reciprocal of the mixedmodel variance (Hedges et al., 1999 ). If the response ratio was not significantly different from unity, growth at elevated [CO2] failed to change leaf O2 uptake from that at current ambient [CO2]. However, if significantly different from 1, growth at elevated [CO2] on average altered O2 uptake across the studies. ACKNOWLEDGMENTS We thank Henry Ginsberg and Joseph Veltre of AEI Technologies, and Nicholas Dowling of Qubit Systems for technical help and modifications to the O2 analysis system; Jeff Amthor, Miquel Gonzalez-Meler, Andrew Leakey, and Shawna Naidu for comments on draft versions of the manuscript; Carlos Pimentel and Kate George for the supply of plant material; and Lisa Ainsworth for advice on the meta-analysis. Received July 20, 2003; returned for revision August 20, 2003; accepted September 25, 2003. FOOTNOTES 1 This work was supported by the U.S. Department of Energy’s Office of Science (to B.E.R.), as were the long-term elevated [CO2] experiments in Florida, North Carolina, and Wisconsin. 2 Present address: Department of Biological Sciences, University of Essex, Colchester, CO4 3SQ, UK. 3 Present address: Dipartimento di Scienze dell’ Ambiente Forestale e delle Sue Risorse, Università della Tuscia, Via S. Camillo de Lellis, 01100 Viterbo, Italy. Article, publication date, and citation information www.plantphysiol.org/cgi/doi/10.1104/pp.103.030569. • can be found at Corresponding author; e-mail stevel@life.uiuc.edu; fax 217-244-7563. LITERATURE CITED Ainsworth EA, Davey PA, Bernacchi CJ, Dermody OC, Heaton EA, Moore DJ, Morgan PB, Naidu SL, Yoo Ra HS, Zhu XG et al. (2002) A meta-analysis of elevated [CO2] effects on soybean (Glycine max) physiology, growth and yield. Glob Change Biol 8: 695709[CrossRef][ISI] Ainsworth EA, Tranel PJ, Drake BG, Long SP (2003) The clonal structure of Quercus geminata revealed by conserved microsatellite loci. Mol Ecol 12: 527532[CrossRef][ISI][Medline] Amthor JS (1997) Plant respiratory responses to elevated CO2 partial pressure. In LH Allen, MB Kirkham, DM Olszyk, CE Whitman, eds, Advances in Carbon Dioxide Effects Research. American Society of Agronomy Special Publication, Madison, WI, pp 35-77 Amthor JS (2000) Direct effect of elevated CO2 on nocturnal in situ leaf respiration in nine temperate deciduous tree species is small. Tree Physiol 20: 139-144[ISI][Medline] Amthor JS, Koch GW, Bloom AJ (1992) CO2 inhibits respiration in leaves of Rumex crispus L. Plant Physiol 98: 757-760[ISI] Amthor JS, Koch GW, Willms JR, Layzell DB (2001) Leaf O2 uptake in the dark is independent of coincident CO2 partial pressure. J Exp Bot 52: 2235-2238[Abstract/Free Full Text] Azcon-Bieto L, Osmond CB (1983) Relationship between photosynthesis and respiration: the effect of carbohydrate status on the rate of CO2 production by respiration in darkened and illuminated wheat leaves. Plant Physiol 71: 574-581[ISI] Baker JT, Allen LH, Boote KJ, Pickering NB (2000) Direct effects of atmospheric carbon dioxide concentration on whole canopy dark respiration of rice. Glob Change Biol 6: 275286[CrossRef][ISI] Bingham GE, Coyne PI (1977) A portable, temperature-controlled, steady-state porometer for field measurements of transpiration and photosynthesis. Photosynthetica 11: 148-160[ISI] Bingham GE, Coyne PI, Kennedy RB, Jackson WL (1980) Design and fabrication of a portable minicuvette system for measuring leaf photosynthesis and stomatal conductance under controlled conditions. UCRL-52895. Lawrence Livermore National Laboratory, Livermore, CA Bloom AJ, Mooney HA, Bjorkman O, Berry JA (1980) Materials and methods for carbon dioxide and water exchange analysis. Plant Cell Environ 3: 371-376[ISI] Bouma TJ, Nielsen KL, Eissenstat DM, Lynch JP (1997) Soil CO2 concentration does not affect growth or root respiration in bean or citrus. Plant Cell Environ 20: 14951505[CrossRef][ISI] Bruhn D, Mikkelsen TN, Atkin OK (2002) Does the direct effect of atmospheric CO2 concentration on leaf respiration vary with temperature? Responses in two species of Plantago that differ in relative growth rate. Physiol Plant 114: 57-64[CrossRef][ISI][Medline] Bunce JA (1995) The effect of carbon dioxide concentration on respiration of growing and mature soybean leaves. Plant Cell Environ 18: 575-581[ISI] Bunce JA (2001) Effects of prolonged darkness on the sensitivity of leaf respiration to carbon dioxide concentration in C3 and C4 species. Ann Bot 87: 463-468[Abstract/Free Full Text] Bunce JA (2002) Carbon dioxide concentration at night affects translocation from soybean leaves. Ann Bot 90: 399-403[Abstract/Free Full Text] Burton AJ, Pregitzer KS (2002) Measurement carbon dioxide concentration does not affect root respiration of nine tree species in the field. Tree Physiol 22: 67-72[ISI][Medline] Cornic G, Jarvis PG (1972) Effects of oxygen on CO2 exchange and stomatal resistance in Sitka spruce and maize at low irradiances. Photosynthetica 6: 225-239[ISI] Curtis PS, Wang X (1998) A meta-analysis of elevated CO2 effects on woody plant mass, form and physiology. Oecologia 113: 299-313[CrossRef][ISI] DeLucia EH, Hamilton JG, Naidu SL, Thomas RB, Andrews JA, Finzi A, Lavine M, Matamala R, Mohan JE, Hendrey GR et al. (1999) Net primary production of a forest ecosystem with experimental CO2 enrichment. Science 284: 1177-1179[Abstract/Free Full Text] Dickson RE, Lewin KF, Isebrands JG, Coleman MD, Heilman WE, Riemenschneider DE, Sober J, Host GE, Hendrey GR, Pregitzer KS et al. (2001) Forest atmosphere carbon transfer storage-II (FACTS II): the aspen free-air CO2 and O3 enrichment (FACE) project in an overview. General Tech. Rep. NC-214. U.S. Department of Agriculture Forest Service North Central Experiment Station, Rhinelander, WI. Drake BG, Azcon-Bieto J, Berry J, Bunce J, Dijkstra J, Farrar J, Gifford RM, GonzàlezMeler MA, Koch G, Lambers H et al. (1999) Does elevated atmospheric CO2 concentration inhibit mitochondrial respiration in green plants? Plant Cell Environ 22: 649-657[CrossRef][ISI] Drake BG, Gonzalez-Meler M, Long SP (1997) More efficient plants: a consequence of rising atmospheric CO2. Annu Rev Plant Physiol Plant Mol Biol 48: 607-637 Gonzalez-Meler MA, Siedow JN (1999) Direct inhibition of mitochondrial respiratory enzymes by elevated CO2: Does it matter at the tissue or whole-plant level? Tree Physiol 19: 253259[ISI][Medline] Griffin KL, Anderson OR, Gastrich MD, Lewis JD, Lin GH, Schuster W, Seemann JR, Tissue DT, Turnbull MH, Whitehead D (2001) Plant growth in elevated CO2 alters mitochondrial number and chloroplast fine structure. Proc Natl Acad Sci USA 98: 24732478[Abstract/Free Full Text] Gurevitch J, Hedges LV (1999) Statistical issues in ecological meta-analyses (meta-analysis in ecology). Ecology 80: 1142-1150[ISI] Hamilton JG, Thomas RB, Delucia EH (2001) Direct and indirect effects of elevated CO2 on leaf respiration in a forest ecosystem. Plant Cell Environ 24: 975-982[CrossRef][ISI] Hedges LV, Gurevitch J, Curtis PS (1999) The meta-analysis of response ratios in experimental ecology. Ecology 80: 1150-1156[ISI] Holtum JAM, Winter K (2003) Photosynthetic CO2 uptake in seedlings of two tropical tree species exposed to oscillating elevated concentrations of CO2. Planta 218: 152158[CrossRef][ISI][Medline] Hunt S (2003) Measurements of photosynthesis and respiration in plants. Physiol Plant 117: 314325[CrossRef][ISI][Medline] Hymus GJ, Snead TG, Johnson DP, Hungate BA, Drake BG (2002) Acclimation of photosynthesis and respiration to elevated atmospheric CO2 in two Scrub Oaks. Glob Change Biol 8: 317-328[CrossRef][ISI] Jach ME, Ceulemans R (2000) Short-versus long-term effects of elevated CO2 on night-time respiration of needles of Scots pine (Pinus sylvestris L.). Photosynthetica 38: 5767[CrossRef][ISI] Jahnke S (2001) Atmospheric CO2 concentration does not directly affect leaf respiration in bean or poplar. Plant Cell Environ 24: 1139-1151[CrossRef][ISI] Jahnke S, Krewitt M (2002) Atmospheric CO2 concentration may directly affect leaf respiration measurement in tobacco, but not respiration itself. Plant Cell Environ 25: 641-651[CrossRef][ISI] Long SP, Farage PK, Garcia RL (1996) Measurement of leaf and canopy photosynthetic CO2 exchange in the field. J Exp Bot 47: 1629-1642[Abstract] Moore BD, Cheng SH, Sims D, Seemann JR (1999) The biochemical and molecular basis for photosynthetic acclimation to elevated atmospheric CO2. Plant Cell Environ 22: 567582[CrossRef][ISI] Morgan PB, Ainsworth EA, Long SP (2003) How does elevated ozone impact soybean? A meta-analysis of photosynthesis, growth and yield. Plant Cell Environ 26: 13171328[CrossRef][ISI] Reuveni J, Gale J, Mayer AM (1993) Reduction of respiration by high ambient CO2 and the resulting error in measurement of respiration made with O2 electrodes. Ann Bot 72: 129131[Abstract/Free Full Text] Robertson EJ, Williams M, Harwood JL, Lindsay JG, Leaver CJ, Leech RM (1995) Mitochondria increase 3-fold and mitochondrial proteins and lipid change dramatically in postmeristematic cells in young wheat leaves grown in elevated CO2. Plant Physiol 108: 469474[Abstract/Free Full Text] Rogers A, Allen DJ, Davey PA, Morgan PB, Ainsworth EA, Bernacchi CJ, Cornic G, Dermody O, Heaton EA, Mahoney J et al. (2004) Leaf photosynthesis and carbohydrate dynamics of soybeans grown throughout their life-cycle under free-air carbon dioxide enrichment. Plant Cell Environ 27 (in press) Rosenburg MS, Adams DC, Gurevitch J (2000) MetaWin: Statistical Software for MetaAnalysis, Version 2.0. Sinauer Associates, Sunderland, MA Teskey RO (1995) A field study of the effects of elevated CO2 on carbon assimilation, stomatal conductance and leaf and branch growth of Pinus taeda trees. Plant Cell Environ 18: 565-573[ISI] Tissue DT, Lewis JD, Wullschleger SD, Amthor JS, Griffin KL, Anderson R (2002) Leaf respiration at different canopy positions in sweetgum (Liquidambar styraciflua) grown in ambient and elevated concentrations of carbon dioxide in the field. Tree Physiol 22: 11571166[ISI][Medline] Tjoelker MG, Oleksyn J, Lee TD, Reich PB (2001) Direct inhibition of leaf dark respiration by elevated CO2 is minor in 12 grassland species. New Phytol 150: 419-424[CrossRef][ISI] von Caemmerer S, Farquhar GD (1981) Some relationships between the biochemistry of photosynthesis and the gas exchange of leaves. Planta 153: 376-387[ISI] Wang XZ, Curtis P (2002) A meta-analytical test of elevated CO2 effects on plant respiration. Plant Ecol 161: 251-261[CrossRef][ISI] Willms JR, Dowling AN, Dong ZM, Hunt S, Shelp BJ, Layzell DB (1997) The simultaneous measurement of low rates of CO2 and O2 exchange in biological systems. Anal Biochem 254: 272-282[CrossRef][ISI][Medline] Willms JR, Salon C, Layzell DB (1999) Evidence for light-stimulated fatty acid synthesis in soybean fruit. Plant Physiol 120: 1117-1127[Abstract/Free Full Text] HOME HELP FEEDBACK ASPB JOURNALS SUBSCRIPTIONS ARCHIVE PLANT PHYSIOLOGY SEARCH TABLE OF CONTENTS THE PLANT CELL Copyright © 2004 by the American Society of Plant Biologists AOBPreview originally Annals of Botany doi:10.1093/aob/mch189 published 2004 online on September 8, 2004 94(5):647-656; This Article Abstract Full Text (PDF) All Versions 94/5/647 mch189v1 of this most Article: recent Alert me when this article is cited Alert me if a correction is posted Services Email this article to a friend Similar articles in this journal Annals of Botany 94/5, © Annals of Botany Company 2004; all rights reserved Similar articles in ISI Web of Science Similar articles in PubMed Download to citation manager INVITED REVIEW PubMed Plant Respiration and Elevated Atmospheric CO2 Concentration: Cellular Responses and Global Significance PubMed Citation Articles by GONZALEZ-MELER, M. A. Articles by TRUEMAN, R. J. MIQUEL A. GONZALEZ-MELER*, TANEVA and REBECCA J. TRUEMAN LINA Agricola Department of Biological Sciences, University of Illinois at Chicago, 845 West Taylor St, Chicago, IL 60607, USA • Articles by GONZALEZ-MELER, M. A. Articles by TRUEMAN, R. J. For correspondence. E-mail mmeler@uic.edu Received: 2 June 2004 September 2004 Returned for revision: 14 June 2004 Accepted: 6 July 2004 Published electronically: 8 TOP ABSTRACT INTRODUCTION DIRECT AND INDIRECT EFFECTS... INTEGRATED EFFECTS OF ELEVATED... ACKNOWLEDGEMENTS LITERATURE CITED ABSTRACT • Background Elevated levels of atmospheric [CO2] are likely to enhance photosynthesis and plant growth, which, in turn, should result in increased specific and whole-plant respiration rates. However, a large body of literature has shown that specific respiration rates of plant tissues are often reduced when plants are exposed to, or grown at, high [CO2] due to direct effects on enzymes and indirect effects derived from changes in the plant’s chemical composition. • Scope Although measurement artefacts may have affected some of the previously reported effects of CO2 on respiration rates, the direction and magnitude for the effects of elevated [CO2] on plant respiration may largely depend on the vertical scale (from enzymes to ecosystems) at which measurements are taken. In this review, the effects of elevated [CO2] from cells to ecosystems are presented within the context of the enzymatic and physiological controls of plant respiration, the role(s) of non-phosphorylating pathways, and possible effects associated with plant size. • Conclusions Contrary to what was previously thought, specific respiration rates are generally not reduced when plants are grown at elevated [CO2]. However, whole ecosystem studies show that canopy respiration does not increase proportionally to increases in biomass in response to elevated [CO2], although a larger proportion of respiration takes place in the root system. Fundamental information is still lacking on how respiration and the processes supported by it are physiologically controlled, thereby preventing sound interpretations of what seem to be species-specific responses of respiration to elevated [CO2]. Therefore the role of plant respiration in augmenting the sink capacity of terrestrial ecosystems is still uncertain. Key words: Respiration, elevated CO2, cellular processes, ecosystem respiration, oxidation INTRODUCTION Respiration is essential for growth and maintenance of all plant tissues, and plays an important role in the carbon balance of individual cells, whole-plants and ecosystems, as well as in the global carbon cycle. Through the processes of respiration, solar energy conserved during photosynthesis and stored as chemical energy in organic molecules is released in a regulated manner for the production of ATP, the universal currency of biological energy transformations, and reducing power (e.g. NADH and NADPH). A quantitatively important by-product of respiration is CO2 and, therefore, plant and ecosystem respiration play a major role in the global carbon cycle. Terrestrial ecosystems exchange about 120 Gt of carbon per year with the atmosphere, through the processes of photosynthesis and respiration (Schlesinger, 1997 ). Roughly, half of the CO2 assimilated annually through photosynthesis is released back to the atmosphere by plant respiration (Gifford, 1994 ; Amthor, 1995 ). Because terrestrial biosphere-atmosphere fluxes of CO2 far outweigh anthropogenic inputs of CO2 to the atmosphere, a small change in terrestrial respiration could have a significant impact on the annual increment in atmospheric [CO2] (Amthor, 1997 ). A large body of literature has indicated that plant respiration is reduced in plants grown at high [CO2]. For example, it was estimated that the observed 15–20 % reduction in plant tissue respiration caused by doubling current atmospheric [CO2] (Amthor, 1997 ; Drake et al., 1997 ; Curtis and Wang, 1998 ), could increase the net sink capacity of global ecosystems by 3·4 Gt of carbon per year (Drake et al., 1999 ), and thus offset an equivalent amount of carbon from anthropogenic CO2 emissions. Therefore, not only are gross changes in respiration important for large-scale carbon balance issues, changes in specific rates of respiration can have significant impact on basic plant biology such as growth, biomass partitioning or nutrient uptake (Amthor, 1991 ; Wullschleger et al., 1994 ; Drake et al., 1999 ). Scaling the effects of an increase in atmospheric [CO2] on plant respiration at the biochemical level to the whole ecosystem is difficult for at least two important reasons: (1) the fine and coarse control points of respiratory pathways in tissues and whole plants are not well known; and (2) it is unclear how respiration rates actively or passively adjust to the effects that elevated [CO2] may have both upstream (e.g. on substrate availability) and downstream (e.g. on energy demand) of the carbon oxidation pathways. In addition, accurate measurements of respiratory rates as CO2 evolution have been proven difficult with current techniques (Gonzalez-Meler and Siedow, 1999 ; Hamilton et al., 2001 ; Jahnke, 2001 ), presenting yet another limitation to scaling up processbased measurements of respiration rates to organism, ecosystem or global levels. In this document, we re-evaluate the theory on respiration responses to elevated [CO2] in view of recent studies that have provided new insights on the effects of a short- and long-term change in atmospheric [CO2] on plant respiration. TOP ABSTRACT INTRODUCTION DIRECT AND INDIRECT EFFECTS... INTEGRATED EFFECTS OF ELEVATED... ACKNOWLEDGEMENTS LITERATURE CITED DIRECT AND INDIRECT EFFECTS OF [CO2] ON PLANT-SPECIFIC RESPIRATION RATES Amthor (1991) described two different effects of [CO2] on plant respiration rates that can be distinguished experimentally: a direct effect, in which the rate of respiration of mitochondria or tissues could be rapidly and reversibly reduced following a rapid increase in [CO2] (Amthor et al., 1992 ; Amthor, 2000a ), and an indirect effect in which high [CO2] changes the rate of respiration in plant tissues compared with the rate seen for those plants grown in normal ambient [CO2] when both are tested at a common background [CO2] (Azcón-Bieto et al., 1994 ). Most of the studies investigating the effects of a change in [CO2] on respiration rates have described magnitude (from non-significant to 60 %) and direction (either stimulation, inhibition or no effect) of these two effects, with little progress on their underlying mechanisms. Confounding measurement artefacts are an added complication for establishing the presence and magnitude of direct and indirect effects of [CO2] on respiration (Gonzalez-Meler and Siedow, 1999 ; Amthor, 2000a ; Jahnke, 2001 ; Davey et al., 2004 ). Direct effects Previously, a rapid short-term doubling of current atmospheric CO2 levels was reported to inhibit respiration of mitochondria and plant tissues by 15–20 % (Amthor, 1997 ; Drake et al., 1997 ; Curtis and Wang, 1998 ). Although direct effects of [CO2] on respiratory enzymes have been reported, the magnitude of the direct inhibition of intact tissue respiration by elevated [CO2] has now been shown to be explained by measurement artefacts (i.e. Gonzalez-Meler and Siedow, 1999 ; Jahnke, 2001 ), diminishing the impact that such reductions in respiration rate may have on plant growth and the global carbon cycle. Direct effects on enzymes Doubling current levels in [CO2] (here we review the effects of only two- or three-fold increases in atmospheric CO2) have been shown to inhibit the oxygen uptake of isolated mitochondria and the activity of mitochondrial enzymes under some conditions (Reuveni et al., 1995 ; GonzàlezMeler et al., 1996a , b ). Twice current atmospheric [CO2] reduces the activity of cytochrome c oxidase and succinate dehydrogenase in isolated mitochondria from cotyledon and roots of Glycine max (Gonzàlez-Meler et al., 1996b ). Mitochondrial oxygen uptake may show either no response or up to 15 % inhibition by a rapid increase in [CO2], depending on the electron donor used in the assay (Gonzàlez-Meler et al., 1996b ; see also Fig. 1), indicating that the response of mitochondrial enzymes to high [CO2] depends on the cell’s metabolism (Gonzalez-Meler and Siedow, 1999 ; Affourtit et al., 2001 ). FIG. 1. Effect of doubling the concentration of CO2 on the oxygen uptake of mitochondria isolated from 5-d-old soybean cotyledons in state 4 conditions (absence of ADP). Oxidation of NADH and pyruvate was measured in intact mitochondria that were incubated for 10 min with 0 (open bars) or 0·1 (solid bars) mM dissolved inorganic carbon (DIC) at 25 °C. Experiments were done using the oxygen isotope technique as described in Ribas-Carbo et al. (1995) to detect the effects of CO2 on the total activity of mitochondrial respiration (Vt), the activity of the View larger version (19K): cytochrome pathway (vcyt) and the activity of the [in this window] alternative pathway (valt). Results are average of four [in a new window] replicates and standard errors. In scaling the direct effects of high [CO2] on respiratory enzymes to intact tissues two factors need to be considered: (1) cytochrome oxidase exerts up to 50 % of the total respiratory control (for definitions, see Kacser and Burns, 1979 ) at the level of the mitochondrial electron transport chain (reviewed by Gonzalez-Meler and Siedow, 1999 ); and (2) the competitive nature of the cytochrome and alternative pathways of mitochondrial respiration (Affourtit et al., 2001 ). The activity of the alternative pathway of respiration could increase upon doubling of the [CO2], masking a direct CO2 inhibition of the cytochrome pathway (Fig. 1). The oxygen isotope technique (Ribas-Carbo et al., 1995 ) allows for the distinction of direct effects of [CO2] on oxygen uptake activity by the cytochrome and the alternative oxidases. Figure 1 shows that mitochondrial electron transport activity is affected when CO2 (a mild inhibitor) restricts the normal electron flow through one of the pathways (Cyt pathway – vcyt). Under conditions of low alternative pathway activity (for details, see Ribas-Carbo et al., 1995 ), inhibition of the cytochrome pathway (18 %) by doubling the ambient [CO2] is compensated for by a similar increase in alternative pathway activity (valt), resulting in no significant reduction in overall oxygen uptake of the isolated mitochondria. These results show that enhanced activity of the alternative pathway upon an increase in [CO2] provides a mechanism by which the direct [CO2] inhibition of the cytochrome oxidase may not be seen in the total oxygen uptake under conditions of high adenylate control (low ADP). Gonzàlez-Meler and Siedow (1999) argued that direct effects of CO2 on respiration exceeding >10 % inhibition are likely due to factors other than inhibition of mitochondrial enzymes. Due to confounding measurement artefacts (Jahnke, 2001 ), rates of tissue respiration have been reported to decrease >10 % when the atmospheric [CO2] doubles. Direct effects on intact tissues Recent evidence indicates that, in studying direct effects of CO2 on respiratory CO2 efflux of intact tissues, investigators should first be concerned about measurement artefacts. Amthor (1997) and Gonzalez-Meler and Siedow (1999) showed that gas CO2 exchange measurement errors could augment and explain the magnitude of the direct effect. New studies made in this context have shown that respiration rates are little or not at all inhibited by a doubling of atmospheric [CO2] (Amthor, 2000a ; Amthor et al., 2001 ; Bunce, 2001 ; Hamilton et al., 2001 ; Tjoelker et al., 2001 ; Bruhn et al., 2002 ; Davey et al., 2004 ). Indeed, Jahnke (2001) and Jahnke and Krewitt (2002) confirmed that measurement artefacts due to leakage in CO2exchange systems could be as large as the previously reported direct inhibitory effects. They also found that leaks through the aerial spaces of homobaric leaves showed a significant apparent inhibition of CO2 efflux that was not due to an inhibition of respiration by elevated [CO2]. Therefore, considerations of the impact direct effects of [CO2] on plant respiration rates may have on the global carbon cycle were overstated (Gonzalez-Meler et al., 1996a ; Drake et al., 1999 ). Corrections for instrument leaks can be applied in most cases (Bunce, 2001 ; Jahnke, 2001 ; Pons and Welschen, 2002 ). Applying some corrections for gas exchange leaks, Bunce (2001) reported a significant reduction in respiration after a rapid increase in [CO2]. Moreover, simultaneously switching the air isotopic composition from 12CO2 to pure 13CO2 when the [CO2] was changed, showed that the efflux of 12CO2 (which represents net respiratory CO2 evolution originating from respiratory substrates minus the CO2 being re-fixed by carboxylases, see below) was reduced in leaves of C3 plants, but not C4, when the [CO2] increased (Pinelli and Loreto, 2003 ). These recent results illustrate that not all reports of direct inhibition of respiration by elevated [CO2] have yet been reconciled with each other, but they do not show a clear proof that the direct effect, as defined by Amthor (1991) exits either. An alternate explanation for the apparent reduction in respiration rates upon an increase in [CO2] (Bunce, 2001 ; Pinelli and Loreto, 2003 ) is an increase in dark CO2 fixation catalysed by phosphoenolpyruvate carboxylase (PEPC), resulting in an apparent reduction of CO2 efflux (Amthor, 1997 ; Drake et al., 1999 ). However, in Rumex and Glycine leaves, Amthor et al. (2001) found an effect of elevated [CO2] on CO2 efflux in only a minority of experiments, with no effect of elevated [CO2] on the O2 uptake of the same leaves, indicating at most a small effect of elevated [CO2] on PEPC activity or respiration rates. Similar results were found for a variety of species, involving 600 measurements, in which [CO2] increases did not alter the CO2 and O2 leaf exchanges in the dark (Davey et al., 2004 ). It seems that direct effects of [CO2] on mitochondrial enzymes may have no consequence on the specific respiratory rate of whole tissues. Because mitochondrial respiratory enzymes are generally ‘in excess’ to the levels required to support normal tissue respiratory activity under most growth conditions (Gonzalez-Meler and Siedow, 1999 ; Atkin and Tjoelker, 2003 ; Gonzalez-Meler and Taneva, 2004 ), a minor inhibition of enzymatic activity by [CO2] will not translate to the overall tissue respiration rate. A small increase in enzyme levels in plants exposed to elevated [CO2] could be enough to compensate for the direct effects of CO2 on mitochondrial enzyme activity, as it has been seen in leaves of some plants grown at high [CO2] (i.e. Griffin et al., 2001 ). Another factor to consider is that increases in alternative pathway activity upon inhibition of cytochrome oxidase by a change in [CO2] will result in unaltered dark respiratory rates (as illustrated in Fig. 1). However, an increase in the non-phosphorylating activity of the alternative pathway reduces the efficiency with which energy from oxidized substrates supports growth and maintenance of plants (Gonzalez-Meler and Siedow, 1999 ; Gonzalez-Meler and Taneva, 2004 ). Some studies have reported that the growth of plants exposed to elevated [CO2] only during the night-time is altered (Bunce, 1995 , 2001 , 2002 ; Reuveni et al., 1997 ; Griffin et al., 1999 ). Such altered growth patterns have been attributed, in part, to effects of [CO2] on plant respiration, including increased activity of the non-phosphorylating alternative pathway. These CO2 effects are considered next. Are there direct effects of CO2 on whole plants? In the past, in some studies, plants have been grown at [CO2] elevated only at night-time to assess the long-term consequences of the previously considered direct effects of CO2 on respiration rates. Although most of these studies have found that growth patterns were altered in plants exposed to high night-time [CO2], these effects cannot be attributed to direct effects of [CO2] on respiration for reasons explained above. Bunce (1995) observed that the biomass and leaf area ratio of Glycine max increased and that photosynthetic rates decreased in plants exposed to high nighttime [CO2] relative to plants grown at normal ambient [CO2]. Similarly, Griffin et al. (1999) found that Glycine max exposed to high night-time [CO2] had lower leaf respiration rates and greater biomass than plants grown at ambient or elevated [CO2]. Reuveni et al. (1997) speculated that increases in the biomass of Lemna gibba grown at high night-time [CO2] relative to control plants, was due to a reduction in alternative pathway respiration (although the alternative pathway activity was not measured). Reduction in alternative pathway activity might more fully couple respiration rates with growth and maintenance, enhancing growth. In contrast, Ziska and Bunce (1999) showed that elevation of night-time [CO2] reduced biomass only in two of the four C4 species studied. In a later study, Bunce (2002) described that carbohydrate translocation was reduced within 2 d of exposing plants to elevated night-time [CO2] when compared with plants grown at ambient conditions day and night. These results suggest that elevated [CO2] may have other uncharacterized direct effects on plant physiology that can have the consequence of reducing energy demand for carbohydrate translocation, hence reducing the rate of leaf respiration. If so, these new types of effects cannot be catalogued as direct effects of [CO2] on respiration but as indirect effects, making more research in this area necessary. In summary, even though rapid changes in [CO2] can inhibit the activity of some mitochondrial enzymes directly, previously reported direct effects of [CO2] on tissue respiration are likely to be due to measurement artefacts. Therefore, direct effects of [CO2] on specific respiration rates (although not necessarily on respiration physiology) should be dismissed as having a major impact on the amount of anthropogenic carbon that vegetation could retain. The role of the alternative pathway in direct respiratory responses to elevated [CO2], although unresolved, would have little impact, if any, on the general conclusion that direct effects of [CO2] on plant respiration are not to be considered in plant growth or carbon cycle models. Other effects of long-term elevation of night-time [CO2] can alter plant growth characteristics, and indirectly affect both respiration rates and the relative activity of the cytochrome and alternative pathways. Therefore these effects cannot be referred to as direct effects but rather as indirect effects of [CO2] on respiration rates. Indirect effects Indirect CO2 effects represent changes in tissue respiration in response to plant growth at elevated atmospheric [CO2] (Amthor, 1997 ), because of changes in tissue composition (Drake et al., 1997 ). Other indirect effects of [CO2] on plant respiration include changes in growth or response to environmental stress, as well as changes in the respiratory demand for energy relative to that of plants grown at ambient [CO2] (Bunce, 1994 ; Amthor, 2000b ). Such indirect effects of plant growth at elevated [CO2] can be detected as a reduction in CO2 emission (or O2 consumption) from plant tissues when measured at ambient [CO2] (Gonzalez-Meler et al., 1996a ). Indirect effects on enzymes: is there acclimation of respiration to elevated [CO2]? Acclimation of respiration to high [CO2] can be defined as the down- or up-regulation of the respiratory machinery (i.e. amount of respiratory enzymes, number of mitochondria) irrespective of changes in specific respiration rates when plants are grown at elevated [CO2] (Drake et al., 1999 ). A change in the leaf respiratory machinery of plants grown at high [CO2] can be expected because of (a) increases in carbohydrate content, (b) reduced photorespiratory activity and (c) reductions in leaf protein content (15 % on average), when compared with plants grown at current ambient [CO2] (Drake et al., 1997 ). These three processes can have different and sometimes opposite effects on levels of respiratory enzymes and specific rates of respiration. As atmospheric [CO2] rises, increased photosynthesis results in higher cellular carbohydrate concentrations (Drake et al., 1997 ; Curtis and Wang, 1998 ). Increased carbohydrates can stimulate the specific activity of respiration, due to the greater availability of respiratory substrates (Azcón-Bieto and Osmond, 1983 ) and higher energy demand for phloem loading of carbohydrates (Bouma et al., 1995 ; Körner et al., 1995 ; Amthor, 2000b ). Additionally, increased tissue carbohydrate levels (as in the case of plants grown at high [CO2]) could result in an increase of transcript levels of cytochrome oxidase (Felitti and Gonzalez, 1998 ; Curi et al., 2003 ) and cytochrome pathway activity (Gonzalez-Meler et al., 2001 ). In contrast, a reduction in photorespiratory activity in plants grown at high [CO2] will reduce the need for the mitochondrial compartment, which may result in a reduction in mitochondrial proteins and functions (Amthor, 1997 ; Drake et al., l999 ; Bloom et al., 2002 ). Not surprisingly, the available studies show that levels and activity of respiratory machinery can either increase or decrease in leaves of plants grown at high [CO2] with no concomitant effects on respiration rates (Fig. 2). FIG. 2. Relationship between the relative response of leaf respiration and the leaf respiratory machinery of plants grown at ambient and elevated [CO2]. Data extracted from Azcon-Bieto et al. (1994) , George et al. (1996) , Gonzalez-Meler et al. (1996a) , Griffin et al. (2001 , 2004) , Hrubec et al. (1985) , Tissue et al. (2002) , Wang et al. (2004) and M. A. Gonzalez-Meler (unpubl. res.). * Respiratory machinery refers to changes in either number of View larger version (14K): mitochondria, or soluble or membrane enzyme activity in [in this window] plant extracts from plants grown at either ambient or [in a new window] elevated [CO2] in field and laboratory experiments. Only studies that have provided both respiration rates and respiratory machinery from the same tissues or from different tissues from the same plants are considered. Rates of [CO2] evolution used are only those measured for plants grown under ambient or elevated [CO2] at ambient [CO2] conditions. For instance, indirect effects of elevated [CO2] on respiration of leaves of Lindera benzoin and stems of Scirpus olneyi were correlated with a reduction in maximum activity of cytochrome oxidase (Azcon-Bieto et al., 1994 ). However, a reduction of respiratory enzyme activity was not seen in rapidly growing tissues exposed to elevated [CO2] (Perez-Trejo, 1981 ; Hrubeck et al., 1985 ). No acclimation to elevated [CO2] (i.e. reduction in cytochrome oxidase) has been observed in leaves of C4 plants (Azcon-Bieto et al., 1994 ) or roots of C3 plants (Gonzalez-Meler et al., 1996a ). In contrast, the number of mitochondria is consistently shown to increase in response to plant growth at elevated [CO2] (Griffin et al., 2001 ; Wang et al., 2004 ). Overall changes associated with increases or decreases in the respiratory machinery in leaves of plants grown at elevated [CO2] (i.e. mitochondrial counts, cytochrome oxidase maximum activity) seem to be independent of responses of the respiration rates to elevated [CO2] (Fig. 2; r2 = 0·, n.s.). Therefore, changes in the mitochondrial machinery of plants grown in high [CO2] are associated with altered mitochondrial function and biogenesis and are not necessarily related to energy production. Anaplerotic functions of mitochondrial activity provide reductant and carbon skeletons for the production of primary and secondary metabolites used in growth and maintenance processes, as well as other processes such as nitrogen reduction and assimilation (Gonzalez-Meler et al., 1996a ; Amthor, 1997 ; Bloom et al., 2002 ; Davey et al., 2004 ). Vanoosten et al., (1992) reported increases in fumarase and malic enzyme activities in leaves of Picea abies grown in high [CO2]. These increases contrasted with unchanged or decreased activity of PEPC or glycolitic enzymes. Increase in enzyme activity of the tricarboxylic acid cycle with no increase in respiration rates may support the export of carbon skeletons and reducing power (mainly as malate) from the mitochondria to the cytosol for biosynthesis. This uncoupling between an increase in mitochondrial numbers and no increase in respiratory activity of plants grown at high [CO2] (Fig. 2) is further illustrated by the study of Tissue et al. (2002) in which a more than two-fold increase in the leaf mitochondrial counts did not alter the maximum activity of cytochrome oxidase in the same tissues. With the exception of two studies (i.e. Azcon-Bieto et al., 1994 ; Tissue et al., 2002 ), there are no reports on the response of membrane-associated mitochondrial enzymes to elevated [CO2] in leaves or roots of plants. More research is needed in this area. Indirect effects of [CO2] on respiration of tissues and whole plants It has been historically accepted that leaf respiration is reduced as a consequence of plant growth at elevated [CO2] (El Kohen et al., 1991 ; Isdo and Kimball, 1992 ; Wullschleger et al., 1992a ; Amthor, 1997 ; Drake et al., 1997 ; Curtis and Wang, 1998 ; Norby et al., 1999 ). Poorter et al. (1992) showed that leaf respiration was reduced, on average, by 14 % when expressed on a leaf mass basis, but increased by 16 % on a leaf area basis. Curtis and Wang (1998) compiled respiratory data for woody plants, and observed that growth at elevated [CO2] resulted in an 18 % inhibition of leaf respiration (mass basis). In this section we will concentrate on the effects of elevated [CO2] on respiration of above-ground tissues. It is worth noting, however, that root respiration rates in the field are not altered by growth at elevated [CO2] during most of the plant’s growing season (Johnson et al., 1994 ; Matamala and Schlesinger, 2000 ). Total ecosystem root respiration may increase if root mass and/or turnover increases (Hungate et al., 1997 ; Hamilton et al., 2002 ; George et al., 2003 ; Matamala et al., 2003 ). Many previous reviews and meta-analyses (see above) have compared respiratory rates of plants grown and measured at the [CO2] at which plants were grown. Therefore, these respiration measurements are also susceptible by the gas exchange artefacts described by Jahnke and coworkers. Respiratory O2 uptake measured in closed systems is not susceptible to the measurement artefacts described above. The O2 uptake of green tissues of plants grown at high [CO2] could be increased, reduced or unaltered when compared with plants grown at ambient [CO2] (Table 1). Davey et al. (2004) also showed that the O2 uptake of leaves of plants grown at high [CO2] was slightly increased relative to control plants. Gas exchange leaks (Jahnke, 2001 ) should not be a factor in determining the rates of CO2 emission rates when rates are measured and compared at ambient [CO2] for plants grown at ambient and elevated [CO2]. A re-analysis of the literature focused on leaf respiratory responses (on a leaf mass basis) of plants grown at ambient and elevated [CO2] and measured at an ambient [CO2], suggests that specific leaf respiration rates will be unaltered or even increased in plants grown at elevated [CO2] (Table 2). Therefore, the generally accepted onclusion that respiration rates of plants grown at elevated [CO2] is reduced relative to plants grown at ambient [CO2] should be re-evaluated (Amthor, 2000a ; Davey et al., 2004 ; Gonzalez-Meler and Taneva, 2004 ). However, there is significant variability in the leaf respiratory response to growth at elevated [CO2] when compared with plants grown at ambient conditions, ranging from 40 % inhibition (Azcon-Bieto et al., 1994 ) to 50 % stimulation (Williams et al., 1992 ). Considerations on the physiological basis by which the acclimation response of respiration to elevated [CO2] varies, is considered next. View this table: [in this window] [in a new TABLE 1. Indirect effects of [CO2] on the dark O2 uptake rates (on a leaf area basis) at 25 °C of green tissues from Scirpus olneyi (C3), Spartina patens (C4), Distchilis spicata (C4), Lindera benzoin (C3), Cinna arundinacea (C3), Quercus geminata (C3), Quercus myrtifolia (C3) and Serenoa repens (C3) plants grown at either normal ambient (A) or normal ambient + 340 mL L–1 atmospheric CO2 window] (E), inside open-top chambers in the field View this TABLE 2. Indirect effects of the elevation in atmospheric [CO2] on the dark CO2 efflux of leaves on a dry mass basis based on literature surveys table: [in this window] [in a new window] In boreal species, reduction of respiration of plants grown at high [CO2] was related to changes in tissue N and carbohydrate concentration (Tjoelker et al., 1999 ). Tissue N concentration often decreases in plants grown at elevated [CO2] (Drake et al., 1997 ; Curtis and Wang, 1998 ). It is expected that respiration rate would be lower in tissues having lower [N], because the respiratory cost associated with protein turnover and maintenance is a large portion of dark respiration (Bouma et al., 1994 ). Hence, the metabolic cost (i.e. respiratory energy demand) for construction and maintenance of tissues with high protein concentration is greater than the cost for the maintenance of the same tissue with low [N] (assuming there are no changes in rates of protein turnover between plants grown at ambient and elevated [CO2]) (Amthor, 1989 ; Drake et al., 1999 ). This idea was confirmed for leaves of Quercus alba seedlings grown in open-top chambers in the field, where respiration was 21–56 % lower under elevated [CO2] than under ambient [CO2] (Wullschleger and Norby, 1992 ). The growth respiration component of these leaves was reduced by 31 % and the maintenance component by 45 %. These effects were attributed to the reduced cost of maintaining tissues having lower [N]. Similar results were obtained in leaves and stems of plants of other species (Wullschleger et al., 1992a , b , 1997 ; Amthor et al., 1994 ; Carey et al., 1996 ; Griffin et al., 1996b ; Dvorak and Oplustilova, 1997 ; Will and Ceulemans, 1997 ), with some exceptions (e.g. Wullschleger et al., 1995 ). The growth component of respiration (Amthor, 2000b ) could also be reduced in plants grown at elevated [CO2] as a result of altered tissue chemistry (Griffin et al., 1993 ). Based on the chemical composition of tissues, Poorter et al. (1997) found that elevated [CO2] could reduce growth construction costs by 10–20 %. Griffin et al. (1993 , 1996a) also observed reductions in construction costs of Pinus taeda seedlings grown at elevated [CO2]. Hamilton et al. (2001) reported that elevated [CO2] slightly reduced construction costs of leaves of mature trees (including P. taeda) at the top of the canopy, but not at the bottom of the canopy. Such a small reduction can be explained by reductions in tissue [N], as observed in leaves exposed to high [CO2] at the top of the canopy. Changes in construction costs did not result in a decrease in the leaf respiration rates of P. taeda trees exposed to elevated [CO2] (Hamilton et al., 2001 ). The lack of long-term effects of increased [CO2] on specific plant respiration rates could also be due to a lower involvement of the alternative pathway (Gonzalez-Meler and Siedow, 1999 ; Griffin et al., 1999 ). Respiration through the alternative pathway bypasses two of the three sites of proton translocation; so the free energy released is lost as heat, and is unavailable for the synthesis of ATP. Respiration associated with this pathway will not support growth and maintenance processes of tissues as efficiently as respiration through the cytochrome path. On the other hand, the activity of the alternative pathway of respiration could decrease upon doubling [CO2], masking increases in the activity of the cytochrome pathway and making respiration more efficient, as it was the case in understorey trees grown under elevated [CO2] (Gonzalez-Meler and Taneva, 2004 ). Despite earlier reports, the responses of plant respiration to growth under high [CO2] are variable and perhaps species specific, although the overall trend may be a moderate increase in respiration rates of leaves of plants grown under elevated [CO2] relative to the ambient ones (Tables 1 and 2). Ultimately, altered specific respiration rates of tissues will depend on the net balance between the demand for ATP from maintenance (including phloem loading and unloading) and growth processes and on carbon allocation patterns between sinks and source tissues. The fact that the mitochondrial machinery has been shown to increase in leaves of plants grown at elevated [CO2] suggests, however, a larger participation of mitochondria in functions other than oxidative phosphorylation. Finally, indirect effects of [CO2] on tissue respiration rates can be augmented or offset by changes in plant size and/or changes in carbon allocation between plant parts. TOP ABSTRACT INTRODUCTION DIRECT AND INDIRECT EFFECTS... INTEGRATED EFFECTS OF ELEVATED... ACKNOWLEDGEMENTS LITERATURE CITED INTEGRATED EFFECTS OF ELEVATED [CO2] ON PLANT RESPIRATION AT THE ECOSYSTEM LEVEL Terrestrial ecosystems exchange about 120 Gt of carbon per year with the atmosphere, through the processes of photosynthesis (leading to gross primary production, GPP) and ecosystem respiration (Re) (Schlesinger, 1997 ). The difference between GPP and Re determines net ecosystem productivity (NEP), the net amount of carbon retained or released by a given ecosystem. Currently, the net exchange of C between the terrestrial biosphere and the atmosphere is estimated to result in a global terrestrial sink of about 2 Gt of carbon per year (Gifford, 1994 ; Schimel, 1995 ; Steffen et al., 1998 ). Unfortunately, the effects of [CO2] on plant and heterotrophic respiration at the ecosystem level are not well understood, despite their potential to control ecosystem carbon budgets (Ryan, 1991 ; Giardina and Ryan, 2000 ; Valentini et al., 2000 ). Annually, terrestrial plant respiration releases 40–60 % of the total carbon fixed during photosynthesis (Gifford, 1994 ; Amthor, 1995 ) representing about half of the annual input of CO2 to the atmosphere from terrestrial ecosystems (Schlesinger, 1997 ). Therefore the magnitude of terrestrial plant respiration and its responses to [CO2] are important factors governing the intrinsic capacity of ecosystems to store carbon. Plant respiration responses to high [CO2] may stem from several mechanisms (in absence of direct effects on respiration rates): (a) indirect effects; (b) changes in total plant biomass; and (c) changes in plant carbon allocation. If it is confirmed that the response of overall terrestrial plant respiration to an increase in atmospheric [CO2] is small (Tables 1 and 2), then changes in global plant respiration should be proportional to changes in biomass. However, experimental evidence suggests otherwise, or that the response of plant respiration at the ecosystem level to elevated [CO2] may be more a function of carbon allocation patterns rather than just of increases in plant size (Drake et al., 1996 ; Hamilton et al., 2002 ). Attempts to scale [CO2] effects on mitochondrial or tissue respiration to the ecosystem level are problematic because, unlike photosynthesis, little is known about applicable scaling methods for plant respiration (Gifford, 2003 ). Current attempts to build respiratory carbon budgets at the canopy level require knowledge of maintenance and growth respiration (see above), as well as tissue respiratory responses to light, temperature and [CO2] (Amthor, 2000b ; Gifford, 2003 ; Turnbull et al., 2003 ) or have been based on theoretical respiration-to-photosynthesis ratios (i.e. Norby et al., 2002 ). The respiration-to-photosynthesis ratio was not affected by elevated [CO2] in soybean plants (Ziska and Bunce, 1998 ) but was reduced in pine forests stands (Hamilton et al., 2002 ) when compared with ambient controls. The limited available data on the effects of elevating the [CO2] on intact field ecosystems on plant and total ecosystem respiration are summarized in Tables 3 and 4. View this TABLE 3. Contribution of CO2-induced reduction in ecosystem respiration (Re) to total net ecosystem productivity in a C3-dominated salt-marsh ecosystem table: [in this exposed to elevated [CO2] since 1988 window] [in a new window] View table: [in window] [in a window] this TABLE 4. The effect of elevated [CO2] on ecosystem-level plant gas exchange of forested ecosystems this new The contribution of CO2-induced changes in ecosystem respiration to annual NEP of an intact salt marsh exposed to elevated [CO2] in open-top chambers is shown in Table 3. In this ecosystem, elevation of atmospheric CO2 by 350 µL L–1 over ambient levels consistently reduced night-time ecosystem respiration in C3 and C4 community stands (Drake et al., 1996 ). Net ecosystem productivity increased in stands exposed to elevated [CO2] because the decrease in respiration was accompanied by increases in GPP. The reduction in annual respiration in this ecosystem represented from one-fifth to one-third of the difference in NEP between ambient and elevated open-top chambers (Table 3). Ecosystem-level plant respiration in forest stands, however, may not be reduced by elevated [CO2] as in the case of the salt marsh (Table 4). In forests exposed to elevated [CO2] using free air CO2 enrichment (FACE), GPP was stimulated by 15–36 % (Table 4) over ambient stands, resulting in consistent increases in net primary productivity (NPP). Increases in GPP and NPP seemed to be accompanied by variable responses on whole-plant respiration, which could be increased by as much as 20 % (Table 4). For instance, Hamilton et al. (2002) found that elevated [CO2] increased forest NPP by 27 % without any increase in total plant ecosystem respiration. This result can be explained by a possible decrease in specific respiration rates at high [CO2] accompanied by larger biomass of plants grown at elevated [CO2]. In a Popolus deltoides plantation at Biosphere 2 centre, elevated [CO2] increased standlevel plant respiration by 40 % when compared with ambient (Table 4). The total increase in stand-level plant respiration was larger than the CO2 stimulation of NPP, suggesting that, in this case, specific respiration rates and plant size both contributed to the increase in plant respiration of trees grown in high [CO2] (R. J. Trueman and M. A. Gonzalez-Meler, unpubl. res.). Some studies suggest that increases in ecosystem-level plant respiration in ecosystems exposed to elevated [CO2] mainly occur in below-ground plant tissues (Hungate et al., 1997 ; Lin et al., 2001 ; Hamilton et al., 2002 ), which in turn may stimulate soil respiration rates (Zak et al., 2000 ; Pendall et al., 2003 ). This requires a substantial proportion of the additional C assimilated by plants growing at elevated [CO2] to be allocated to roots for growth and turnover (Johnson et al., 1994 ; Hungate et al., 1997 ; King et al., 2001 ; Matamala et al., 2003 ). Greater plant C allocation below ground can theoretically increase the contribution of root respiration to total soil respiration because of greater root biomass relative to ambient [CO2]. However, Hamilton et al. (2002) reported increases in total root respiration in a forest exposed to elevated [CO2], where root mass, turnover (Matamala et al., 2003 ) or respiration rates (Matamala and Schlesinger, 2000 ; George et al., 2003 ) were not affected by the CO2 treatment. The apparent disparity may be due to the different methodologies employed in these studies and in the contribution of rhizosphere activity to soil CO2 efflux, emphasizing the need for more coordinated research in building ecosystem carbon budgets using multiple approaches, especially below ground. In conclusion and contrary to what was previously thought, specific respiration rates are generally not reduced when plants are grown at elevated [CO2]. This is because direct effects of [CO2] on respiratory enzymes are inconsequential to specific respiration rates of tissues. A re-analysis of the literature comparing respiration of leaves of plants grown at ambient [CO2] with leaves of plants grown at elevated [CO2] when rates are measured at the same [CO2], indicate that leaf respiration on average will not be greatly changed by increasing atmospheric [CO2]. Increases in growth rates appear to be compensated for by changes in tissue chemistry that affect growth and maintenance respiration. If specific rates of respiration are not affected by growth at elevated [CO2], respiration from the terrestrial vegetation in a high-CO2 world should be proportional to changes in plant biomass. However, whole ecosystem studies show that canopy respiration does not increase proportionally to increases in biomass when natural ecosystems are exposed to elevated atmospheric [CO2]. Field studies also suggest that a larger proportion of plant respiration takes place in the root system under elevated [CO2] conditions. Fundamental information is still lacking on how respiration and the processes supported by it are physiologically controlled, thereby preventing sound interpretations of what seem to be species-specific responses of respiration to elevated [CO2]. Therefore the role of plant respiration in augmenting the sink capacity of terrestrial ecosystems is still uncertain. TOP ABSTRACT INTRODUCTION DIRECT AND INDIRECT EFFECTS... INTEGRATED EFFECTS OF ELEVATED... ACKNOWLEDGEMENTS LITERATURE CITED ACKNOWLEDGEMENTS This review is part of a lecture given at the International Congress of Plant Mitochondria held in Perth in July 2002, partly sponsored by Annals of Botany. M.A.G.-M. specially thanks Jeff Amthor for his work in this area, his thoughts, discussions and specific comments in parts of this manuscript. Authors also wish to thank Joaquim Azcon-Bieto, Bert Drake, Evan DeLucia, Steve Long, Miquel Ribas-Carbo and Jim Siedow for their discussion on the effects of elevated [CO2] on respiration of organelles, plants and ecosystems over the years. We acknowledge financial support by US Department of Agriculture (M.A.G.-M.), UIC fellowship (L.T.), and Sigma Xi and provost awards to L.T. and R.J.T. LITERATURE CITED Affourtit C, Krab K, Moore AL. 2001. Control of plant mitochondrial respiration. Biochimica Biophysica Acta – Bioenergetics 1504: 58– 69.[CrossRef][ISI] TOP ABSTRACT INTRODUCTION DIRECT AND INDIRECT EFFECTS... INTEGRATED EFFECTS OF ELEVATED... ACKNOWLEDGEMENTS LITERATURE CITED Amthor JS. 1989. Respiration and crop productivity. New York: Springer Verlag. Amthor JS. 1991. Respiration in a future, higher CO2 world. Plant, Cell and Environment 14: 13–20.[ISI] Amthor JS. 1995. Terrestrial higher-plant response to increasing atmospheric [CO2] in relation to the global carbon cycle. Global Change Biology 1: 243–274.[ISI] Amthor JS. 1997. Plant respiratory responses to elevated CO2 partial pressure. In: Allen LH, Kirkham MB, Olszyk DM, Whitman CE, eds. Advances in carbon dioxide effects research. American Society of Agronomy Special Publication (proceedings of 1993 ASA Symposium, Cincinnati, OH), Madison, WI: ASA, CSSA and SSSA, 35–77. Amthor JS. 2000a. Direct effect of elevated CO2 on nocturnal in situ leaf respiration in nine temperate deciduous trees species is small. Tree Physiology 20: 139– 144.[ISI][Medline] Amthor JS. 2000b. The McCree–de Wit–Penning de Vries–Thornley respiration paradigms: 30 years later. Annals of Botany 86: 1–20.[Abstract/Free Full Text] Amthor JS, Koch G, Boom AJ. 1992. CO2 inhibits respiration in leaves of Rumex crispus L. Plant Physiology 98: 1–4.[ISI] Amthor JS, Koch GW, Willms JR, Layzell DB. 2001. Leaf O2 uptake in the dark is independent of coincident CO2 partial pressure. Journal Experimental Botany 52: 2235– 2238. Amthor JS, Mitchell RJ, Runion GB, Rogers HH, Prior SA, Wood CW 1994. Energy content, construction cost and phytomass accumulation of Glycine max (L.) Merr. and Sorghum bicolor (L.) Moench grown in elevated CO2 in the field. New Phytologist 128: 443–450.[ISI] Atkin OK, Tjoelker MG. 2003. Thermal acclimation and the dynamic response of plant respiration to temperature. Trends in Plant Science 8: 343–351.[CrossRef][ISI][Medline] Azcón-Bieto J, Osmond CB. 1983. Relationship between photosynthesis and respiration. The effect of carbohydrate status on the rate of CO2 production by respiration in darkened and illuminated wheat leaves. Plant Physiology 71: 574–581.[ISI] Azcón-Bieto J, Gonzàlez-Meler MA, Doherty W, Drake BG. 1994. Acclimation of respiratory O2 uptake in green tissues of field grown native species after long-term exposure to elevated atmospheric CO2. Plant Physiology 106: 1163–1168.[Abstract/Free Full Text] Bloom AJ, Smart DR, Nguyen DT, Searles PS. 2002. Nitrogen assimilation and growth of wheat under elevated carbon dioxide. Proceedings of the National Academy of Sciences of the USA 99: 1730–1735.[Abstract/Free Full Text] Bouma TJ, De Viser R, Janseen JHJA, De Kick MJ, Van Leeuwen PH, Lambers H. 1994. Respiratory energy requirements and rate of protein turnover in vivo determined by the use of an inhibitor of protein synthesis and a probe to assess its effect. Physiologia Plantarum 92: 585–594.[CrossRef][ISI] Bouma TJ, De Viser R, Van Leeuwen PH, De Kick MJ, Lambers H. 1995. The respiratory energy requirements involved in nocturnal carbohydrate export from starchstoring mature source leaves and their contribution to leaf dark respiration. Journal of Experimental Botany 46: 1185–1194.[ISI] Bruhn D, Mikkelsen TN, Atkin OK. 2002. Does the direct effect of atmospheric CO2 concentration on leaf respiration vary with temperature? Responses in two species of Plantago that differ in relative growth rate. Physiologia Plantarum 114: 57– 64.[CrossRef][ISI][Medline] Bunce JA. 1994. Responses of respiration to increasing atmospheric carbon dioxide concentrations. Physiologia Plantarum 90: 427–430.[CrossRef][ISI] Bunce JA. 1995. Effects of elevated carbon dioxide concentration in the dark on the growth of soybean seedlings. Annals of Botany 75: 365–368.[Abstract/Free Full Text] Bunce JA. 2001. Effects of prolonged darkness on the sensitivity of leaf respiration to carbon dioxide concentration in C3 and C4 species. Annals of Botany 87: 463– 468.[Abstract/Free Full Text] Bunce JA. 2002. Carbon dioxide concentration at night affects translocation from soybean leaves. Annals of Botany 90: 399–403.[Abstract/Free Full Text] Carey EV, DeLucia EH, Ball JT. 1996. Stem maintenance and construction respiration in Pinus ponderosa grown in different concentrations of atmospheric CO2. Tree Physiology 16: 125–130.[ISI] Curi GC, Welchen E, Chan RL, Gonzalez DH. 2003. Nuclear and mitochondrial genes encoding cytochrome c oxidase subunits respond differently to the same metabolic factors. Plant Physiology and Biochemistry 41: 689–693.[CrossRef][ISI] Curtis PS, Wang X. 1998. A meta-analysis of elevated CO2 effects on woody plant mass, form, and physiology. Oecologia 113: 299–313.[CrossRef][ISI] Davey PA, Hunt S, Hymus GJ, DeLucia EH, Drake BG, Karnosky DF, Long SP. 2004. Respiratory oxygen uptake is not decreased by an instantaneous elevation of [CO2], but is increased with long-term growth in the field at elevated [CO2]. Plant Physiology 134: 520–527.[Abstract/Free Full Text] Drake BG, Azcón-Bieto J, Berry JA, Bunce J, Dijkstra P, Farrar J, Koch GW, Gifford R, Gonzàlez-Meler MA, Lambers H, et al. 1999. Does elevated CO2 inhibit plant mitochondrial respiration in green plants? Plant, Cell and Environment 22: 649– 657.[CrossRef][ISI] Drake BG, Gonzalez-Meler MA, Long SP. 1997. More efficient plants: a consequence of rising atmospheric CO2? Annual Review of Plant Physiology and Plant Molecular Biology 48: 609–639.[CrossRef][ISI] Drake BG, Meuhe M, Peresta G, Gonzàlez-Meler MA, Matamala R. 1996. Acclimation of photosynthesis, respiration and ecosystem carbon flux of a wetland on Chesapeake Bay, Maryland, to elevated atmospheric CO2 concentration. Plant and Soil 187: 111–118.[ISI] Dvorak V, Oplustilova M. 1997. Respiration of woody tissues of Norway spruce in elevated CO2 concentrations. In: Mohren GMJ, Kramer K, Sabate S, eds. Impacts of global change on tree physiology and forest ecosystems. Dordrecht: Klurer Academic Publishers, 47–51. El Kohen A, Pontailler Y, Mousseau M. 1991. Effect d’un doublement du CO2 atmosphérique sur la respiration à l’obscurité des parties aeriennes de jeunes chataigniers (Castanea sativa Mill.). Compte rendue de l’Academie des sciences, serie III 312: 477– 481. Felitti SA, Gonzalez DH. 1998. Carbohydrates modulate the expression of the sunflower cytochrome c gene at the mRNA level. Planta 206: 410–415.[CrossRef][ISI] George V, Gerant D, Dizengremel P. 1996. Photosynthesis, Rubisco activity and mitochondrial malate oxidation in pedunculate oak (Quercus robur L.) seedlings grown under present and elevated atmospheric CO2 concentrations. Annales des Sciences Forestieres 53: 469–474.[ISI] George K, Norby RJ, Hamilton JG, DeLucia EH. 2003. Fine-root respiration in a loblolly pine and sweetgum forest growing in elevated CO2. New Phytologist 160: 511– 522.[CrossRef][ISI] Giardina CP, Ryan MG. 2000. Evidence that decomposition rates of organic carbon in mineral soil do not vary with temperature. Nature 404: 858–861.[CrossRef][ISI][Medline] Gifford RM. 1994. The global carbon cycle: a view point on the missing sink. Australian Journal of Plant Physiology 21: 1–15.[ISI] Gifford RM. 2003. Plant respiration in productivity models: conceptualization, representation and issues for global terrestrial carbon-cycle research. Functional Plant Biology 30: 171–86.[CrossRef][ISI] Gonzàlez-Meler MA, Siedow JN. 1999. Inhibition of respiratory enzymes by elevated CO2: does it matter at the intact tissue and whole plant levels? Tree Physiology 19: 253– 259.[ISI][Medline] Gonzàlez-Meler MA, Taneva L. 2004. Integrated effects of atmospheric CO2 concentration on plant and ecosystem respiration. In: Lambers H, Ribas-Carbo M, eds. Plant respiration. Dordrecht: Kluwer Academic Publishers, in press. Gonzàlez-Meler MA, Drake BG, Azcón-Bieto J. 1996ª. Rising atmospheric carbon dioxide and plant respiration. In: Breymeyer AI, Hall DO, Melillo JM, Ågren GI, eds. Global change: effects on coniferous forests and grasslands. SCOPE, Vol. 56. UK: John Wiley & Sons, 161–181. Gonzàlez-Meler MA, Giles L, RB Thomas, Siedow JN. 2001. Metabolic regulation of leaf respiration and alternative pathway activity in response to phosphate supply. Plant, Cell and Environment 24: 205–215.[CrossRef][ISI] Gonzàlez-Meler MA, Ribas-Carbo M, Siedow JN, Drake BG. 1996b. Direct inhibition of plant mitochondrial respiration by elevated CO2. Plant Physiology 112: 1349– 1355.[Abstract/Free Full Text] Griffin KL, Anderson OR, Gastrich MD, Lewis JD, Lin G, Schuster W, Seemann JR, Tissue DT, Turnbull M, Whitehead D. 2001. Plant growth in elevated CO2 alters mitochondrial number and chloroplast fine structure. Proceedings of the National Academy of Science of the USA 98: 2473–2478.[CrossRef] Griffin KL, Anderson OR, Tissue DT, Turnbull MH, Whitehead D. 2004. Variations in dark respiration and mitochondrial numbers within needles of Pinus radiata grown in ambient or elevated CO2 partial pressure. Tree Physiology 24: 347–353.[ISI][Medline] Griffin KL, Ball JT, Strain BR. 1996a. Direct and indirect effects of elevated CO2 on whole-shoot respiration in ponderosa pine seedlings. Tree Physiology 16: 33–41.[ISI] Griffin KL, Sims DA, Seemann JR. 1999. Altered night-time CO2 concentration affects growth, physiology and biochemistry of soybean. Plant, Cell and Environment 22: 91– 99.[CrossRef][ISI] Griffin KL, Thomas RB, Strain BR. 1993. Effects of nitrogen supply and elevated carbon dioxide on construction cost in leaves of Pinus taeda (L.) seedlings. Oecologia 95: 575–580[ISI] Griffin KL, Winner WE, Strain BR. 1996b. Construction cost of loblolly and ponderosa pine leaves grown with varying carbon and nitrogen availability. Plant, Cell and Environment 19: 729–738.[ISI] Hamilton JG, DeLucia EH, George K, Naidu S, Finzi AC, Schlesinger WH. 2002. Forest carbon balance under CO2. Oecologia 131: 250–260.[CrossRef][ISI] Hamilton JG, Thomas RB, Delucia EH. 2001. Direct and indirect effects of elevated CO2 on leaf respiration in a forest ecosystem. Plant, Cell and Environment 24: 975– 982.[CrossRef][ISI] Hrubec TC, Robinson JM, Donaldson RP. 1985. Effects of CO2 enrichment and carbohydrate content on the dark respiration of soybeans. Plant Physiology 79: 684– 689.[ISI] Hungate BA, Holland EA, Jackson RB, Chapin III FS, Mooney HA, Field CB. 1997. The fate of carbon in grasslands under carbon dioxide enrichment. Nature 388: 576– 579.[CrossRef][ISI] Isdo SB, Kimball BA. 1992. Effects of atmospheric CO2 enrichment on photosynthesis, respiration and growth of sour orange trees. Plant Physiology 99: 341–343.[ISI] Jahnke S. 2001. Atmospheric CO2 concentration does not directly affect leaf respiration in bean or poplar. Plant, Cell and Environment 24: 1139–1151.[CrossRef][ISI] Jahnke S, Krewitt M. 2002. Atmospheric CO2 concentration may directly affect leaf respiration measurement in tobacco, but not respiration itself. Plant, Cell and Environment 25: 641–651.[CrossRef][ISI] Johnson DW, Geisinger D, Walker R, Newman J, Vose JM, Elliott KJ, Ball T. 1994. Soil pCO2, soil respiration, and root activity in CO2-fumigated and nitrogen-fertilized ponderosa pine. Plant and Soil 165: 129–138.[ISI] Kacser H, Burns JA. 1979. Molecular democracy: who shares control? Biochemical Society Transactions 7: 1149–1160.[Medline] King JS, Pregitzer KS, Zak DR, Sober J, Isebrands JG, Dickson RE, Hendrey GR, Karnosky DF. 2001. Fine-root biomass and fluxes of soil carbon in young stands of paper birch and trembling aspen as affected by elevated atmospheric CO2 and tropospheric O3. Oecologia 128: 237–250.[CrossRef][ISI] Körner CH, Pelaez-Riedl S, Van Bel AJE. 1995. CO2 responsiveness of plants: a possible link to phloem loading. Plant, Cell and Environment 18: 595–600.[ISI] Lin G, Rygiewicz PT, Ehleringer JR, Johnson MG, Tingey DT. 2001. Time-dependent responses of soil CO2 efflux components to elevated atmospheric [CO2] and temperature in experimental forest mesocosms. Plant and Soil 229: 259–270.[CrossRef][ISI] Matamala R, Schlesinger WH. 2000. Effects of elevated atmospheric CO2 on fine-root production and activity in an intact temperate forest ecosystem. Global Change Biology 6: 967–979.[CrossRef][ISI] Matamala R, Gonzalez-Meler MA, Jastrow JD, Norby RJ, Schlesinger WH. 2003. Impacts of fine root turnover on forest NPP and soil C sequestration potential. Science 302: 1385–1387.[Abstract/Free Full Text] Norby RJ, Hanson PJ, O’Neill EG, Tschaplinski TJ, Weltzin JF, Hansen RA, Cheng WX, Wullschleger SD, Gunderson CA, Edwards NT, et al. 2002. Net primary productivity of a CO2-enriched deciduous forest and the implications for carbon storage. Ecological Applications 12: 1261–1266.[ISI] Norby RJ, Wullschleger SD, Gunderson CA, Johnson DW, Ceulemans R. 1999. Tree responses to rising CO2 in field experiments: implications for the future forest. Plant, Cell and Environment 22: 683–714.[CrossRef][ISI] Pendall E, Del Grosso S, King JY, LeCain DR, Milchunas DG, Morgan JA, Mosier AR, Ojima DS, Parton WA, Tans PP, et al. 2003. Elevated atmospheric CO2 effects and soil water feedbacks on soil respiration components in a Colorado grassland. Global Biogeochemical Cycles 17: 1046. Perez-Trejo MS. 1981. Mobilization of respiratory metabolism in potato tubers by carbon dioxide. Plant Physiology 67: 514–517.[ISI] Pinelli P, Loreto F. 2003. (CO2)-C-12 emission from different metabolic pathways measured in illuminated and darkened C-3 and C-4 leaves at low, atmospheric and elevated CO2 concentration. Journal of Experimental Botany 54: 1761– 1769.[Abstract/Free Full Text] Pons TL, Welschen RAM. 2002. Overestimation of respiration rates in commercially available clamp-on leaf chambers. Complications with measurement of net photosynthesis. Plant, Cell and Environment 25: 1367.[CrossRef][ISI] Poorter H, Gifford RM, Kriedemann PE, Wong SC. 1992. A quantitative analysis of dark respiration and carbon content as factors in the growth response of plants to elevated CO2. Australian Journal of Botany 40: 501–513.[ISI] Poorter H, VanBerkel Y, Baxter R, DenHertog J, Dijkstra P, Gifford RM, Griffin KL, Roumet C, Roy J, Wong SC. 1997. The effect of elevated CO2 on the chemical composition and construction costs of leaves of 27 C-3 species. Plant, Cell and Environment 20: 472–482.[ISI] Reuveni J, Gale J, Mayer AM. 1995. High ambient carbon dioxide does not affect respiration by suppressing the alternative cyanide-resistant respiration. Annals of Botany 76: 291–295.[Abstract/Free Full Text] Reuveni J, Gale J, Zeroni M. 1997. Differentiating day from night effects of high ambient [CO2] on the gas exchange and growth of Xanthium strumarium L. exposed to salinity stress. Annals of Botany 80: 539–546.[Abstract/Free Full Text] Ribas-Carbo M, Berry JA, Yakir D, Giles L, Robinson SA, Lennon AM, Siedow JN. 1995. Electron partitioning between the cytochrome and alternative pathways in plant mitochondria. Plant Physiology 109: 829–837.[Abstract/Free Full Text] Ryan MG. 1991. Effects of climate change on plant respiration. Ecological Applications 1: 157–167.[ISI] Schafer, KVR, Oren R, Ellsworth DS, Lai C, Herricks JD, Finzi AC, Richter DD, Katul GG. 2003. Exposure to an enriched CO2 atmosphere alters carbon assimilation and allocation in a pine forest ecosystem. Global Change Biology 9: 1378– 1400.[CrossRef][ISI] Schimel DS. 1995. Terrestrial ecosystems and the carbon cycle. Global Change Biology 1: 77–91.[ISI] Schlesinger WH. 1997. Biogeochemistry: an analysis of global change, 2nd edn. San Diego, CA: Academic Press. Steffen W, Noble I, Canadell J, Apps M, Schulze ED, Jarvis PG, Baldocchi D, Ciais P, Cramer W, Ehleringer J, et al. 1998. The terrestrial carbon cycle: implications for the Kyoto protocol. Science 280: 1393–1394.[Free Full Text] Tissue DT, Lewis JD, Wullschleger SD, Amthor JS, Griffin KL, Anderson R. 2002. Leaf respiration at different canopy positions in sweetgum (Liquidambar styraciflua) grown in ambient and elevated concentrations of carbon dioxide in the field. Tree Physiology 22: 1157–1166.[ISI][Medline] Tjoelker MG, Oleksyn J, Lee TD, Reich PB. 2001. Direct inhibition of leaf dark respiration by elevated CO2 is minor in 12 grassland species. New Phytologist 150: 419– 424.[CrossRef][ISI] Tjoelker MG, Reich PB, Oleksyn J. 1999. Changes in leaf nitrogen and carbohydrates underlie temperature and CO2 acclimation of dark respiration in five boreal species. Plant, Cell and Environment 22: 767–778.[CrossRef][ISI] Turnbull MH, Whitehead D, Tissue DT, Schuster WSF, Brown KJ, Griffin KL. 2003. Scaling foliar respiration in two contrasting forest canopies. Functional Ecology 17: 101– 114.[CrossRef][ISI] Valentini R, Matteucci G, Dolman AJ, Schulze ED, Rebmann C, Moors EJ, Granier A, Gross P, Jensen NO, Pilegaard K, et al. 2000. Respiration as the main determinant of carbon balance in European forests. Nature 404: 861–865.[CrossRef][ISI][Medline] Vanoosten JJ, Afif D, Dizengremel P. 1992. Long-term effects of a CO2 enriched atmosphere on enzymes of the primary carbon metabolism of spruce trees. Plant Physiology and Biochemistry 30: 541–547.[ISI] Wang XZ, Anderson OR, Griffin KL. 2004. Chloroplast numbers, mitochondrion numbers and carbon assimilation physiology of Nicotiana sylvestris as affected by CO2 concentration. Environmental and Experimental Botany 51: 21–31.[CrossRef][ISI] Will RE, Ceulemans R. 1997. Effects of elevated CO2 concentrations on photosynthesis, respiration and carbohydrate status of coppice Populus hybrids. Physiologia Plantarum 100: 933–939.[CrossRef][ISI] Williams ML, Jones DG, Baxter R, Farrar JF. 1992. The effect of enhanced concentrations of atmospheric CO2 on leaf respiration. In: Lambers H, van der Plas LHW, eds. Molecular, biochemical and physiological aspects of plant respiration. The Hague: SPB Academic Publishing, 547–551. Wullschleger SD, Norby RJ. 1992. Respiratory cost of leaf growth and maintenance in white oak saplings exposed to atmospheric CO2 enrichment. Canadian Journal of Forest Research 22: 1717–1721.[ISI] Wullschleger SD, Norby RJ, Gunderson CA. 1992b. Growth and maintenance respiration in leaves of Liriodendron tulipifera L. exposed to long-term carbon dioxide enrichment in the field. New Phytologist 121: 515–523.[ISI] Wullschleger SD, Norby RJ, Hanson PJ. 1995. Growth and maintenance respiration in stems of Quercus alba after four years of CO2 enrichment. Physiologia Plantarum 93: 47– 54.[CrossRef][ISI] Wullschleger SD, Norby RJ, Hendrix DL 1992a. Carbon exchange rates, chlorophyll content, and carbohydrate status of two forest tree species exposed to carbon dioxide enrichment. Tree Physiology 10: 21–31.[ISI][Medline] Wullschleger SD, Norby RJ, Love JC, Runck C. 1997. Energetic costs of tissue construction in yellow-poplar and white oak trees exposed to long-term CO2 enrichment. Annals of Botany 80: 289–297.[Abstract/Free Full Text] Wullschleger SD, Ziska LH, Bunce JA. 1994. Respiratory response of higher plants to atmospheric CO2 enrichment. Physiologia Plantarum 90: 221–229.[CrossRef][ISI] Zak DR, Pregitzer KS, King JS, Holmes WE. 2000. Elevated atmospheric CO2, fine roots and the response of soil microorganisms: a review and hypothesis. New Phytologist 147: 201–222.[CrossRef][ISI] Ziska LH, Bunce JA. 1998. The influence of increasing growth temperature and CO2 concentration on the ratio of respiration to photosynthesis in soybean seedlings. Global Change Biology 4: 637–643.[CrossRef][ISI] Ziska, LH, Bunce JA. 1999. Effects of elevated carbon dioxide concentration at night on the growth and gas exchange of selected C4 species. Australian Journal of Plant Physiology 26: 71–77.[ISI] This Article Abstract Full Text (PDF) All Versions 94/5/647 mch189v1 of this most Article: recent Alert me when this article is cited Alert me if a correction is posted Services Email this article to a friend Similar articles in this journal Similar articles in ISI Web of Science Similar articles in PubMed Download to citation manager PubMed PubMed Citation Articles by GONZALEZ-MELER, M. A. Articles by TRUEMAN, R. J. Agricola Articles by GONZALEZ-MELER, M. A. Articles by TRUEMAN, R. J. HOME HELP FEEDBACK SUBSCRIPTIONS ARCHIVE SEARCH TABLE OF CONTENTS Plant Physiol, October 2001, Vol. 127, pp. 607-614 Insertional Mutants of Chlamydomonas reinhardtii That Require Elevated CO2 for Survival1 Kyujung Van, Yingjun Wang, Yoshiko Nakamura, and Martin H. Spalding* Interdepartmental Plant Physiology Major (K.V., Y.W., M.H.S.) and Department of Botany (K.V., Y.W., Y.N., M.H.S.), 353 Bessey Hall, Iowa State University, Ames, Iowa 50011 TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED ABSTRACT Aquatic photosynthetic organisms live in quite variable conditions of CO2 availability. To survive in limiting CO2 conditions, Chlamydomonas reinhardtii and other microalgae show adaptive changes, such as induction of a CO2-concentrating mechanism, changes in cell organization, increased photorespiratory enzyme activity, induction of periplasmic carbonic anhydrase and specific polypeptides (mitochondrial carbonic anhydrases and putative chloroplast carrier proteins), and transient down-regulation in the synthesis of Rubisco. The signal for acclimation to limiting CO2 in C. reinhardtii is unidentified, and it is not known how they sense a change of CO2 level. The limiting CO2 signals must be transduced into the changes in gene expression observed during acclimation, so mutational analyses should be helpful for investigating the signal transduction pathway for low CO2 acclimation. Eight independently isolated mutants of C. reinhardtii that require high CO2 for photoautotrophic growth were tested by complementation group analysis. These mutants are likely to be defective in some aspects of the acclimation to low CO2 because they differ from wild type in their growth and in the expression patterns of five low CO2-inducible genes (Cah1, Mca1, Mca2, Ccp1, and Ccp2). Two of the new mutants formed a single complementation group along with the previously described mutant cia-5, which appears to be defective in the signal transduction pathway for low CO2 acclimation. The other mutations represent six additional, independent complementation groups. TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED INTRODUCTION Acclimation to changed environmental conditions is a key to survival for all organisms. In response to perceived environmental signals, organisms may exhibit specific adaptive changes, such as changes in the expression of key genes to survive specific environmental changes. Because CO2 can vary substantially in aquatic habitats and represents the major substrate for photosynthetic CO2 fixation via the enzyme Rubisco, CO2 concentration is an important environmental signal in aquatic photosynthetic organisms including cyanobacteria and Chlamydomonas reinhardtii. Unlike terrestrial higher plants, aquatic photosynthetic organisms can face difficulties in acquiring CO2. Because the CO2 diffusion rate in water is much slower than that in air (Badger and Spalding, 2000 ), the CO2 supply to Rubisco in these aquatic photosynthetic organisms can become limited. C. reinhardtii and other aquatic photosynthetic organisms have a genetic program to allow them to acclimate to low CO2. This acclimation includes induction of a CO2concentrating mechanism (CCM) that allows the cells to acquire CO2 efficiently by increasing the CO2 concentration around Rubisco under limiting CO2 conditions (Badger et al., 1980 ; for review, see Spalding, 1998 ; Kaplan and Reinhold, 1999 ). Along with the induction of the CCM, C. reinhardtii shows adaptive changes to limiting CO2 conditions, such as changes in cell organization (Geraghty and Spalding, 1996 ), increased photorespiratory enzyme activity (Marek and Spalding, 1991 ), induction of periplasmic carbonic anhydrase (CA) (pCA1, encoded by the Cah1 gene; Fujiwara et al., 1990 ; Fukuzawa et al., 1990 ; Ishida et al., 1993 ), mitochondrial CA (mtCA, encoded by the Mca1 and Mca2 genes; Eriksson et al., 1996 ; Geraghty and Spalding, 1996 ), and putative chloroplast carrier protein (Ccp, encoded by the Ccp1 and Ccp2 genes; Geraghty et al., 1990 ; Ramazanov et al., 1993 ; Chen et al., 1997 ), and transient down-regulation in the synthesis of Rubisco (Coleman and Grossman, 1984 ; Winder et al., 1992 ). The signal for acclimation to limiting CO2 in C. reinhardtii is unidentified. It is not known how they sense a change of CO2 availability, whether by CO2 concentration directly or indirectly via a cellular process such as carbohydrate metabolism. Whatever the limiting-CO2 signal, it must be transduced into the changes in gene expression observed during acclimation, such as expression of Cah1. A powerful way to identify components of the CCM and of the signal transduction pathway for low CO2 acclimation is through the analysis and characterization of mutants specifically defective in growth in limiting CO2, like the ca-1, pmp-1, and cia-5 mutants (Spalding et al., 1983a ; 1983b ; Moroney et al., 1989 ). Using advances in nuclear transformation of C. reinhardtii (Kindle, 1990 ), a collection of insertionally generated high CO2-requiring (HCR) mutants unable to grow in limiting CO2 was obtained and is described here. TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED RESULTS Generation and Isolation of Mutants Using glass bead transformation (Kindle, 1990 ; Davies et al., 1994 ), CC425 (Table I) was complemented by transformation with p-Arg7.8 (Debuchy et al., 1989 ) to generate a pool of insertional mutants on CO2-minimal medium. Cells from each of more than 7,000 transformant colonies were suspended in air-minimal medium and grown on plates in high CO2 (5% [v/v] CO2 in air), normal air, and low CO2 (50-100 µL L 1 CO2). HCR mutants, defined as those showing little or no growth either in normal air or in low CO2, should include mutants, like cia-5, that are defective in acclimation to limiting CO2, as well as those with functional defects in the CCM. Sixteen putative HCR mutants were identified, and eight of those are described here (Table II). Table I. Strains of C. reinhardtii used in the study View this table: [in this window] [in a new window] Table II. Characteristics of HCR mutants View this table: [in this window] [in a new window] General Characteristics of HCR Mutants The eight HCR mutants and their general characteristics are shown in Table II and Figure 1. When grown in high CO2 on agar, all HCR mutants except HCR105 were indistinguishable from the wild type (Fig. 1). The eight HCR mutants could be divided into four groups based on their apparent high CO2 requirement for photoautotrophic growth. The first group, including HCRP34, HCR209, and HCR90, showed a leaky HCR phenotype in air but a stringent phenotype in low CO2. The second group, including HCR86 and HCR105, showed a stringent HCR phenotype both in air and in low CO2. HCR89 and HCR95, comprising the third group, had a leaky HCR phenotype both in air and in low CO2. HCR3510 lacked a significant growth phenotype in air but had a stringent phenotype in low CO2. Figure 1. Spot tests for growth response to different CO2 concentrations for wild-type strains (CC849 and ars301), four previously described HCR mutants (cia-5, ca-1, pmp1, and pgp-1), and eight new HCR mutants. Plates were kept either at high CO2 (5% [v/v] CO2), at air level of CO2, or at low CO2 (50-100 µL L 1) for 10 d. View larger version (72K): [in this window] [in a new window] Genetic Characteristics of HCR Mutants Seven of the eight HCR mutants were found by Southern analysis (data not shown) to contain only one copy of the Arg7 insert, and the presence of vector sequences was confirmed in six mutants (Table II). The presence of vector sequences provides an opportunity for the cloning of sequences flanking the insert by plasmid rescue (Quarmby and Hartzell, 1994 ). Selected random progeny and/or tetrads from HCR mutants were tested in crosses with another arg2 mutant (CC1068, Table I) for linkage of the Arg insert with Arg+ and HCR phenotypes (Table II). Five of the eight mutants showed cosegregation of the single Arg insert with the HCR phenotype, suggesting that the Arg insert is responsible for the HCR phenotype in these five mutants. In two of the mutants, HCR95 and HCR105, the inserts did not cosegregate with the HCR phenotype, indicating that insertion of the Arg plasmid was not directly responsible for the HCR phenotype in these two mutants. In HCR209, which has two inserts, cosegregation crosses were not conclusive, but other evidence (see below) suggests the two inserts are tandemly arranged and are responsible for the phenotype. Heterozygous vegetative diploids, generated in crosses with CC1068 and selected by their resistance to both kanamycin and streptomycin, were used to determine the dominant/recessive nature of the HCR phenotype of each mutant. Based on growth tests of the heterozygous diploids, the mutant phenotype of all eight HCR mutants was judged to be recessive. Complementation Group Analysis Crossing with the various known mutants such as cia-5, ca-1, pmp-1, and pgp-1 should help identify new alleles of previously characterized mutants. If any wild-type colonies appear under low CO2 conditions (50-100 µL L 1 CO2) after mating with HCR mutants, this indicates they are not allelic to each other, because these known mutants also show HCR phenotypes. Rapid allelism tests were used to place the various HCR mutants into different complementation groups. Complementation analysis was tested with the eight HCR mutants (Table II) along with cia-5, ca-1, pmp-1, and pgp-1 (Table I). Only crosses between cia-5 × HCRP34, cia-5 × HCR209, and HCRP34 × HCR209 failed to generate wild type colonies. Thus, HCR3510, HCR86, HCR89, HCR90, HCR95, and HCR105 each define a new HCR locus. HCRP34 and HCR209 have been confirmed as defective in the same locus as cia-5 by comparison of the sequence of the DNA flanking the inserts with a cloned cia-5 gene (Xiang et al., 2001 ) and by complementation with a cloned cia-5 gene (data not shown). Liquid Growth Experiments Growth experiments showed patterns of high CO2 requirement for photoautotrophic growth consistent with those seen in spot tests (Fig. 1). Active, 1-d-old air-adapted cells were inoculated into liquid minimal medium with similar starting cell densities (5 × 104 cells ml 1), grown with no aeration, and the cell densities measured daily at the same time of day for 10 d. HCRP34 and HCR209, judged to be allelic to cia-5, grew very similar to cia-5 in air (Fig. 2A). The growth rates of HCR86 and HCR90 also were only slightly better than that of cia-5 in air (Fig. 2B), but the growth rates of HCR89 and HCR95 were intermediate between wild type (ars301; see Table I) and cia-5 (Fig. 2C). HCR105 was able to grow slightly in air but bleached within a few days (Fig. 2B). HCR3510, which showed a wild-type phenotype in air on agar, also grew as well as wild type (ars301) in air in liquid culture (Fig. 2C). Chlorophyll content also was measured in these cultures along with cell density, and the growth curves based on chlorophyll content showed the same pattern as those of cell density (data not shown). Figure 2. Liquid cell growth curves for wild type (ars301), cia-5, and HCR mutants grown at pH 7 on an orbital shaker without aeration. A, HCRP34 and HCR209. B, HCR86, HCR90, and HCR105. C, HCR89, HCR95, and HCR3510. The growth curves shown are averages of three independent growth experiments. View larger version (9K): [in this window] [in a new window] Accumulation of Low CO2-Inducible Transcripts Because the expression of low CO2-inducible polypeptides (pCA1, mtCA1, mtCA2, Ccp1, and Ccp2) has been reported to change differentially during acclimation to limiting CO2 (Villarejo et al., 1996 , 1997 ; Eriksson et al., 1998 ), accumulation of these three transcripts also was analyzed. The cia-5-like mutants, HCRP34 and HCR209, showed no detectable Cah1 mRNA, Mca1 and Mca2 mRNA, and Ccp1 and Ccp2 mRNA (Fig. 3A; data shown only for HCRP34). HCR90, which showed a leaky HCR phenotype in air but a stringent phenotype in low CO2, had reduced expression of only Mca1 and Mca2 mRNA (Fig. 3B). In separate, long-term experiments, the expression of the other genes was somewhat variable, but only Mca1 and Mca2 showed reproducibly decreased mRNA abundance (data not shown). HCR3510, which showed a wildtype phenotype in air but a stringent HCR phenotype in low CO2, had normal expression of these genes compared with wild type (CC849; see Table I; Fig. 3A). However, HCR95 showed a much different pattern of expression for these three genes. From cells exposed for 2 h to air, Cah1 mRNA of HCR95 was detected at normal levels, whereas much-reduced levels of Mca1 and Mca2 mRNA and Ccp1 and Ccp2 mRNA were detected relative to wild type (Fig. 3B). After 6 h, wild type showed the same or increased levels of these three mRNAs, but expression of all three mRNA in HCR95 was dramatically reduced (Fig. 3B), suggesting only a transient induction of their expression in this mutant. In separate, long-term experiments, this apparent transient induction in HCR95 also was confirmed up to 24 h (data not shown). The other HCR mutants (HCR86, HCR89, and HCR105) did not show reproducibly different patterns of expression for the three low CO2-inducible transcripts relative to wild type (data not shown). Figure 3. Northern-blot analyses for wild type (CC849) and HCR mutants. A, HCRP34 and HCR3510. B, HCR90 and HCR95. Total RNA (10 µg per lane) was isolated 2 h to 6 h after transfer of cells to air levels of CO2 from high CO2. Cah1 mRNA was probed with the 1.4-kb BglII and NcoI fragment of Cah1 cDNA (Van and Spalding, 1999 ). Mca1 and Mca2 mRNA was probed with the full-length Mca2 cDNA (Eriksson et al., 1996 , 1998 ). Ccp1 and Ccp2 mRNA was probed with the 1.2-kb EcoRI and HindIII fragment of Ccp1 G1 (Chen et al., 1997 ). The rRNA was probed with 25S and 5.8S rDNA (Marco and Rochaix, View larger version (105K): 1980 ). [in this window] [in a new window] TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED DISCUSSION HCR mutants have been useful for investigation of various processes, both in algae and in higher plants. HCR mutants with defects in several of the enzymes of the photorespiratory pathway have been isolated in the C3 plants Arabidopsis (Somerville and Ogren, 1982 ) and barley (Hordeum vulgare) (Joy et al., 1992 ; Leegood et al., 1996 ; Wingler et al., 1999 ). These photorespiratory mutants exhibited lethality (HCR phenotype) in air levels of CO2 for various reasons, including accumulation of toxic intermediates during photorespiration and depletion of exchangeable nitrogen in photorespiratory intermediates. In C. reinhardtii, the photorespiratory mutant pgp-1 (lacks PGPase) has a HCR phenotype, indicating that the oxygenase activity of Rubisco was not completely suppressed by operation of the CCM and that photorespiratory mutants in C. reinhardtii also are lethal in air levels of CO2 (Suzuki et al., 1990 ; Spalding, 1998 ). Mutants defective in functional components of the CCM also exhibit an HCR phenotype in C. reinhardtii (Spalding et al., 1983a , 1983b ; Moroney et al., 1986 ; Suzuki and Spalding, 1989 ; Funke et al., 1997 ; Karlsson et al., 1998 ) and cyanobacteria (Price and Badger, 1989 ; Ogawa, 1991 , 1992 ; Marco et al., 1993 ; Ohkawa et al., 1998 ; Price et al., 1998 ). Isolation and characterization of the C. reinhardtii mutants, ca-1 and pmp-1, demonstrated the requirement for active transport and accumulation of Ci (Badger et al., 1980 ; Spalding et al., 1983b ) and for a thylakoid lumen CA (Spalding et al., 1983a ; Funke et al., 1997 ; Karlsson et al., 1998 ) for function of the CCM. Another C. reinhardtii HCR mutant, cia-5, exhibits no apparent low-CO2 acclimation responses, such as induction of CCM, up-regulation of low CO2-inducible polypeptides, up-regulation of photorespiratory enzymes, or down-regulation of Rubisco biosynthesis (Moroney et al., 1989 ; Marek and Spalding, 1991 ; Spalding et al., 1991 ; Burow et al., 1996 ). This mutant is thought to be defective in the signal transduction pathway for acclimation to limiting CO2. The gene responsible for this mutation (Cia5) has been cloned recently (Fukuzawa et al., 2001 ; Xiang et al., 2001 ), and its characterization suggests it may encode a transcription factor. The identification of this important gene opens the way for more rapid progress in delineation of the signal transduction pathway for acclimation to limiting CO2. Because many changes involved in acclimation to limiting CO2 conditions appear to be controlled at different gene expression levels, it is possible that mutations in several different loci might yield signal transduction mutants like cia-5 with HCR phenotypes. Thus, the HCR phenotype should be a good indicator of nonacclimation to low CO2 as well as for a dysfunctional CCM, so isolation of HCR mutants should be helpful for identification of loci required for either function of the CCM or for signal transduction leading to low CO2 acclimation. Among the eight new HCR mutants described here, six represent new complementation groups and the other two represent new alleles of the previously described cia-5 locus. The patterns of growth and of low CO2-inducible transcript accumulation for HCRP34 and HCR209 were similar to those of cia-5, and complementation group analyses confirmed that the three are allelic. As new alleles of cia-5, HCRP34 and HCR209 may prove valuable in understanding the function of the gene product from this important locus. Other than for HCRP34 and HCR209, the growth responses to air and low CO2 varied among these new HCR mutants, as did the pattern of accumulation of limiting-CO2-inducible genes. HCR90, which showed a stringent HCR phenotype in low CO2 and grew only slightly better than cia-5 in air (Fig. 2B), had reproducibly reduced expression of only one pair of the limiting-CO2inducible transcripts, Mca1 and Mca2. No disruption of the structural gene for either Mca1 or Mca2 was found in genomic Southern blots probed with the Mca1 and Mca2 promoter region (data not shown), so HCR90 may be defective in a regulatory component that preferentially affects expression of Mca1 and Mca2. HCR86, which has a growth phenotype very similar to HCR90, showed limiting-CO2-inducible transcripts accumulations that were not reproducibly different from those of wild type (data not shown). The leaky phenotype in low CO2 of HCR89 and HCR95 was supported by their growth patterns (Fig. 2C), but only HCR95 reproducibly showed reduced level of low CO2-inducible transcripts (Fig. 3B). HCR3510 showed no significant differences from wild type in terms of low CO2-inducible transcript accumulation, suggesting it is unlikely to be defective in the limiting-CO2-responsive signal transduction pathway. The growth phenotype of this mutant, near wild-type growth in normal air but a stringent phenotype in low CO2, suggests a defect in a functional component of the CCM (or another pathway required for acclimation to limiting CO2) that is essential in very low CO2 but not in air levels of CO2. The advantage of using insertional mutagenesis to generate mutants lies in the use of the inserted DNA as a “tag” to clone the disrupted gene, but of course this only works if the insert cosegregates with the mutant phenotype, i.e. if the insert is responsible for the mutation. As judged by the Arg+ phenotype, the Arg7 inserts in mutants HCRP34, HCR3510, HCR86, HCR89, and HCR90 cosegregate with the HCR phenotype (Table II), suggesting the insert caused the mutation in each of these strains. As indicated above, both HCRP34 and HCR209 are allelic to cia-5 and the insert in each has been confirmed to disrupt the cia-5 gene. Thus, we know the defect in both these mutants, even though cosegregation of the Arg+ and HCR phenotypes has not been demonstrated for HCR209. It is unfortunate that the inserts in mutants HCR95 and HCR105 do not cosegregate with the HCR phenotype, so identification of the disrupted gene responsible for the HCR phenotype in these mutants will have to be accomplished without the aid of an insertional tag. The three remaining tagged mutants (HCR3510, HCR86, and HCR 90) remain as viable candidates for identification of novel genes essential for acclimation of C. reinhardtii to limiting CO2. Cloning of the disrupted genes in these three HCR mutants is in progress. TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED MATERIALS AND METHODS Cell Strains and Culture Conditions All Chlamydomonas reinhardtii strains (Table I) were grown as previously described (Geraghty et al., 1990 ). Cells were cultured on an orbital shaker under aeration with 5% (v/v) CO2 in air (high CO2-grown cells) or no aeration (air-adapted cells). For experiments monitoring the accumulation of low CO2 inducible transcripts, cell cultures were switched from aeration with 5% (v/v) CO2 to aeration with normal air for 2 h to 6 h. For growth on solid media, cells were maintained under 5% (v/v) CO2 in air (high CO2), normal air, or 50 to 100 µL L 1 CO2 (low CO2). Generation and Isolation of Mutants Glass bead transformations were performed as described previously (Van and Spalding, 1999 ). To generate a pool of insertional mutants on CO2-minimal medium, CC425 (Table I) was transformed with linearized p-Arg7.8 (Debuchy et al., 1989 ) containing the structural gene (Arg7) for argininosuccinate lyase to complement the arg2 mutation. Each of more than 7,000 colonies was screened by spot tests to identify HCR mutants. After replica plates with transformants were made, each plate was placed in high CO2 and air or high CO2 and low CO2. Mutants identified in this primary screen as having HCR phenotypes were screened again by western immunoblots of extracellular protein to identify mutants in which pCA1 expression was decreased or absent (Van and Spalding, 1999 ). Spot Growth Tests and Growth Experiments For spot growth tests, actively growing cells were suspended to similar cell densities in minimal medium, spotted (10 µL) onto minimal agar plates, and grown in different concentrations of CO2 for 10 d (Harris, 1989 ). For liquid growth experiments, active, 1-d-old air-adapted cells were inoculated into liquid minimal medium at similar cell densities (5 × 104 cells ml 1). The cultures were grown on an orbital shaker without aeration for the next 10 d. The cell density was determined using a hemacytometer (Reichert Scientific Instruments, Buffalo, NY; Harris, 1989 ). Chlorophyll content was estimated after extraction with 96% (v/v) ethanol (Wintermans and De Mots, 1965 ). DNA- and RNA-Blot Analysis Southern- and northern-blot analyses were performed as described by Van and Spalding (1999) . Total RNA was purified with TRIzol reagent (Life Technologies, Gaithersburg, MD) from airinduced cells exposed to limiting CO2 (aeration with normal air) and Hybond N+ nylon transfer membrane (Amersham Pharmacia Biotech Inc., Piscataway, NJ) was used for blotting. After phoporimager analysis of each northern blot (Molecular Dynamics, Piscataway, NJ), total RNA amounts were normalized to hybridization with 25S and 5.8S rRNA (Marco and Rochaix, 1980 ) using ImageQuaNT (Molecular Dynamics). Genetic Analyses All matings were performed by crossing insertionally generated mutants with various strains (Table I) according to the protocol of Harris (1989) . To isolate vegetative diploids, gametes from HCR mutants (sr-u-2-60) and CC1068 (nr-u-2-1) were induced under nitrogen stress, mated, and the mating mixture spread onto kanamycin-containing medium to select for expression of the plastid-encoded kanamycin resistance (nr-u-2-1) transmitted from the mating-type minus parent. Putative diploids (surviving colonies) were verified by selection for simultaneous expression of the plastid-encoded streptomycin resistance (sr-u-2-60) from the mating-type plus parent and by DNA quantity in flow cytometry (performed at the Iowa State University Cell Facility, Ames). Complementation group analyses required construction of mating type minus strains of each HCR mutant (both new and previously described mutants). Mating type minus strains of cia-5, ca-1, pmp-1, and pgp-1 were generated by crossing with CC124 (Table I). CC1068 (Table I) was used for generating mating type minus strains from all new HCR mutants, except HCR95 and HCR105. After crossing each of the seven new HCR mutants and the four known mutants with each other, the progeny from each cross were tested for photoautotrophic growth in low CO2 (50-100 µL L 1 ). Because all HCR mutants required elevated CO2 for survival, wild-type colonies were observed in low CO2 only if the cross generated wild-type recombinant progeny. FOOTNOTES Received April 9, 2001; returned for revision June 4, 2001; accepted July 2, 2001. 1 This work was supported by the U.S. Department of Agriculture National Research Initiative (grant nos. 97-35100-4210 and 99-35100-7569 to M.H.S.). This is journal paper no. J-19297 of project no. 3578 of the Iowa Agriculture and Home Economics Experiment Station (Ames) and was supported by the Hatch Act and State of Iowa funds. • Corresponding author; e-mail mspaldin@iastate.edu; fax 515-294-1377. Article, publication date, and citation www.plantphysiol.org/cgi/doi/10.1104/pp.010333. information can be found at TOP ABSTRACT INTRODUCTION RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED LITERATURE CITED Badger MR, Kaplan A, Berry JA (1980) Internal inorganic carbon pool of Chlamydomonas reinhardtii: evidence for a CO2 concentrating mechanism. Plant Physiol 66: 407-413[ISI] Badger MR, Spalding MS (2000) CO2 acquisition, concentration and fixation in cyanobacteria and algae. In RC Leegood, TD Sharkey, S von Caemmerer, eds, Photosynthesis: Physiology and Metabolism. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp 369-397 Burow MD, Chen Z-Y, Mouton TM, Moroney JV (1996) Isolation of cDNA clones induced upon transfer of Chlamydomonas reinhardtii cells to low CO2. Plant Mol Biol 31: 443-448[ISI][Medline] Chen Z-Y, Lavigne MD, Mason CB, Moroney JV (1997) Cloning and overexpressing of two cDNAs encoding the low-CO2-inducible chloroplast envelope protein LIP-36 from Chlamydomonas reinhardtii. Plant Physiol 114: 265-273[Abstract/Free Full Text] Coleman JR, Grossman AR (1984) Biosynthesis of carbonic anhydrase in Chlamydomonas reinhardtii during adaptation to low CO2. Proc Natl Acad Sci USA 81: 6049-6053[Abstract] Davies JP, Yildiz F, Grossman AR (1994) Mutants of Chlamydomonas with aberrant responses to sulfur deprivation. Plant Cell 6: 53-63[Abstract/Free Full Text] Debuchy R, Purton S, Rochaix JD (1989) The argininosuccinate lyase gene of Chlamydomonas reinhardtii: an important tool for nuclear transformation and for correlating the genetic and molecular maps of the ARG7 locus. EMBO J 8: 28032809[Abstract] Eriksson M, Karlsson J, Ramazanov Z, Garderstrom P, Samuelsson G (1996) Discovery of an algal mitochondrial carbonic anhydrase: molecular cloning and characterization of low-CO2-induced polypeptide in Chlamydomonas reinhardtii. Proc Natl Acad Sci USA 93: 12031-12034[Abstract/Free Full Text] Eriksson M, Villand P, Garderstrom P, Samuelsson G (1998) Induction and regulation of expression of a low-CO2-induced mitochondrial carbonic anhydrase in Chlamydomonas reinhardtii. Plant Physiol 116: 637-641[Abstract/Free Full Text] Fujiwara S, Fukuzawa H, Tachiki A, Miyachi S (1990) Structure and differential expression of two genes encoding carbonic anhydrase in Chlamydomonas reinhardtii. Proc Natl Acad Sci USA 87: 9779-9783[Abstract] Fukuzawa H, Fujiwara S, Yamamoto Y, Dionisio-Sese ML, Miyachi S (1990) cDNA cloning, sequence, and expression of carbonic anhydrase in Chlamydomonas reinhardtii: regulation by environmental CO2 concentration. Proc Natl Acad Sci USA 87: 43834387[Abstract] Fukuzawa H, Miura K, Ishizaki K, Kucho KI, Saito T, Kohinata T, Ohyama K (2001) Ccm1, a regulatory gene controlling the induction of a carbon-concentrating mechanism in Chlamydomonas reinhardtii by sensing CO2 availability. Proc Natl Acad Sci USA 98: 5347-5352[Abstract/Free Full Text] Funke RP, Kovar JL, Weeks DP (1997) Intracellular carbonic anhydrase is essential to photosynthesis in Chlamydomonas reinhardtii at atmospheric levels of CO2. Plant Physiol 114: 237-244[Abstract/Free Full Text] Geraghty AM, Anderson JC, Spalding MH (1990) A 36 kilodalton limiting-CO2 induced polypeptide of Chlamydomonas is distinct from the 37 kilodalton periplasmic carbonic anhydrase. Plant Physiol 93: 116-121[ISI] Geraghty AM, Spalding MH (1996) Molecular and structural changes in Chlamydomonas under limiting CO2: a possible mitochondrial role in adaptation. Plant Physiol 111: 13391347[Abstract/Free Full Text] Harris EH (1989) The Chlamydomonas Source Book: A Comprehensive Guide to Biology and Laboratory Use. Academic Press, San Diego Ishida S, Muto S, Miyachi S (1993) Structural analysis of periplasmic carbonic anhydrase 1 of Chlamydomonas reinhardtii. Eur J Biochem 214: 9-16[Abstract] Joy KW, Blackwell RD, Lea PJ (1992) Assimilation of nitrogen in mutants lacking enzymes of the glutamate synthase cycle. J Exp Bot 43: 139-145[ISI] Kaplan A, Reinhold L (1999) CO2 concentrating mechanisms in photosynthetic microorganisms. Ann Rev Plant Physiol Mol Biol 50: 539-570[CrossRef][ISI] Karlsson J, Clarke AK, Chen ZY, Hugghins SY, Park YI, Husic HD, Moroney JV, Samuelsson G (1998) A novel alpha-type carbonic anhydrase associated with the thylakoid membrane in Chlamydomonas reinhardtii is required for growth at ambient CO2. EMBO J 17: 1208-1216[Abstract/Free Full Text] Kindle KL (1990) High-frequency nuclear transformation of Chlamydomonas reinhardtii. Proc Natl Acad Sci USA 87: 1228-1232[Abstract] Leegood RC, Lea PJ, Hausler RE (1996) Use of barley mutants to study the control of photorespiratory metabolism. Biochem Soc Trans 24: 757-761[ISI][Medline] Marco E, Ohad N, Schwarz R, Lieman Hurwitz J, Gabay C, Kaplan A (1993) High CO2 concentration alleviates the block in photosynthetic electron transport in an ndhBinactivated mutant of Synechococcus sp. PCC7942. Plant Physiol 101: 10471053[Abstract/Free Full Text] Marco Y, Rochaix JD (1980) Organization of the nuclear ribosomal DNA of Chlamydomonas reinhardtii. Mol Gen Genet 177: 715-723[ISI][Medline] Marek LF, Spalding MH (1991) Changes in photorespiratory enzyme activity in response to limiting CO2 in Chlamydomonas reinhardtii. Plant Physiol 97: 420-425[ISI] Moroney JV, Husic HD, Tolbert NE, Kitayama M, Manuel LJ, Togasaki RK (1989) Isolation and characterization of a mutant of Chlamydomonas reinhardtii deficient in the CO2 concentration mechanism. Plant Physiol 89: 897-903[ISI] Moroney JV, Tolbert NE, Sears BB (1986) Complementation analysis of the inorganic carbon concentration mechanism of Chlamydomonas reinhardtii. Mol Gen Genet 204: 199203[ISI] Ogawa T (1991) Cloning and inactivation of a gene essential to inorganic carbon transport of Synechocystis PCC6803. Plant Physiol 96: 280-284[ISI] Ogawa T (1992) Identification and characterization of the ictA/ndhL gene product essential to inorganic carbon transport of Synechocystis PCC6803. Plant Physiol 99: 1604-1608[ISI] Ohkawa H, Sonoda M, Katoh H, Ogawa T (1998) The use of mutants in the analysis of the CCM in cyanobacteria. Can J Bot 76: 1035-1042[CrossRef][ISI] Price GD, Badger MR (1989) Isolation and characterization of the high CO2-requiringmutants of the cyanobacterium Synechococcus PCC7942. Plant Physiol 91: 514-525[ISI] Price GD, Sültemeyer D, Klughammer B, Ludwig M, Badger MR (1998) The functioning of the CO2 concentrating mechanism in several cyanobacterial strains: a review of general physiological characteristics, genes, proteins and recent advances. Can J Bot 76: 973-1002[CrossRef][ISI] Quarmby LM, Hartzell HC (1994) Dissection of eukaryotic transmembrane signaling using Chlamydomonas. Trends Pharmacol Sci 15: 343-349[CrossRef][ISI][Medline] Ramazanov Z, Mason CB, Geraghty AM, Spalding MH, Moroney JV (1993) The low CO2-inducible 36 kD protein is localized to the chloroplast envelope of Chlamydomonas reinhardtii. Plant Physiol 101: 1195-1199[Abstract/Free Full Text] Somerville CR, Ogren WL (1982) Genetic modification of photorespiration. Trends Biochem Sci 7: 171-174[CrossRef][ISI] Spalding MH (1998) CO2 acquisition: adaptation to changing carbon availability. In J-D Rochaix, M Goldschmidt-Clermont, S Merchant, eds, The Molecular Biology of Chloroplasts and Mitochondria in Chlamydomonas. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp 529-547 Spalding MH, Spreitzer RJ, Ogren WL (1983a) Carbonic anhydrase deficient mutant of Chlamydomonas requires elevated carbon dioxide concentration for photoautotrophic growth. Plant Physiol 73: 268-272[ISI] Spalding MH, Spreitzer RJ, Ogren WL (1983b) Reduced inorganic carbonic transport in a CO2-requiring mutant of Chlamydomonas reinhardtii. Plant Physiol 73: 273-276[ISI] © Spalding MH, Winder TL, Anderson JC, Geraghty AM, Marek LF (1991) Changes in protein and gene expression during induction of the CO2-concentrating mechanism in wildtype and mutant Chlamydomonas. Can J Bot 69: 1008-1016[ISI] Suzuki K, Marek LF, Spalding MH (1990) A photorespiratory mutant of Chlamydomonas reinhardtii. Plant Physiol 93: 92-96 Suzuki K, Spalding MH (1989) Adaptation of Chlamydomonas reinhardtii high-CO2requiring mutants to limiting-CO2. Plant Physiol 90: 1195-1200[ISI] Van K, Spalding MH (1999) Periplasmic carbonic anhydrase structural gene (Cah1) mutant in Chlamydomonas reinhardtii. Plant Physiol 120: 757-764[Abstract/Free Full Text] Villarejo A, Garcia Reina G, Ramazanov Z (1996) Regulation of the low-CO2-inducible polypeptides in Chlamydomonas reinhardtii. Planta 199: 481-485[ISI] Villarejo A, Martinez F, Ramazanov Z (1997) Effect of aminooxyacetate, an inhibitor blocking the glycolate pathway, on the induction of a CO2-concentrating mechanism and low-CO2-inducible polypeptides in Chlamydomonas reinhardtii (Chlorophyta). Eur J Phycol 32: 141-145[CrossRef][ISI] Winder TL, Anderson JC, Spalding MH (1992) Translational regulation of the large and small subunits of ribulose bisphosphate carboxylase/oxygenase during induction of the CO2-concentration mechanism in Chlamydomonas reinhardtii. Plant Physiol 98: 14091414[ISI] Wingler A, Quick WP, Bungard RA, Bailey KJ, Lea PJ, Leegood RC (1999) The role of photorespiration during drought stress: an analysis utilizing barley mutants with reduced activities of photorespiratory enzymes. Plant Cell Environ 22: 361-373[CrossRef][ISI] Wintermans JFGM, De Mots A (1965) Spectrophotometric characteristics of chlorophyll and their pheophytins in ethanol. Biochem Biophys Acta 109: 448-453[ISI][Medline] Xiang Y, Zhang J, Weeks DP (2001) The Cia5 gene controls formation of the carbon concentrating mechanism in Chlamydomonas reinhardtii. Proc Natl Acad Sci USA 98: 5341-5346[Abstract/Free Full Text] 2001 American Society of Plant Physiologists This article has been cited by other articles: M. C. Posewitz, S. L. Smolinski, S. Kanakagiri, A. Melis, M. Seibert, and M. L. Ghirardi Hydrogen Photoproduction Is Attenuated by Disruption of an Isoamylase Gene in Chlamydomonas reinhardtii PLANT CELL, August 1, 2004; 16(8): 2151 2163. [Abstract] [Full Text] [PDF] This Article Abstract Full Text (PDF) Alert me when this article is cited Alert me if a correction is posted Services Similar articles in this journal Similar articles in ISI Web of Science Similar articles in PubMed Alert me to new issues of the journal Download to citation manager Search for citing articles ISI Web of Science (3) in: PubMed PubMed Citation Articles by Van, K. Articles by Spalding, M. H. Agricola Articles by Van, K. Articles by Spalding, M. H. HOME HELP FEEDBACK SUBSCRIPTIONS ASPB JOURNALS ARCHIVE SEARCH TABLE OF CONTENTS PLANT PHYSIOLOGY THE PLANT CELL Copyright © 2001 by the American Society of Plant Biologists Plant First Physiology published on December 30, Preview 2003; Plant Physiology, January 2004, Vol. 134, pp. 520-527 This Article 10.1104/pp.103.030569 ENVIRONMENTAL STRESS AND ADAPTATION Respiratory Oxygen Uptake Is Not Decreased by an Instantaneous Elevation of [CO2], But Is Increased with Long-Term Growth in the Field at Elevated [CO2]1 Abstract Full Text (PDF) All Versions of this 134/1/520 most pp.103.030569v1 Article: recent Alert me when this article is cited Phillip A. Davey2, Stephen Hunt, Graham J. Hymus3, Evan H. DeLucia, Bert G. Drake, David F. Karnosky and Stephen P. Long* Departments of Crop Sciences and Plant Biology, University of Illinois, Urbana, Illinois 61801 (P.A.D., S.P.L., E.H.D.); Department of Biology, Queen’s University, Kingston, Ontario, K7L 3N6, Canada (S.H.); Smithsonian Environmental Research Center, Edgewater, Maryland 21307 (G.J.H., B.R.D.); and School of Forest Resources and Environmental Science, Michigan Technological University, Houghton, Michigan 49931 (D.F.K.) Alert me if a correction is posted Services Similar articles in this journal Similar articles in ISI Web of Science Similar articles in PubMed Alert me to new issues of the journal Download to citation manager Cited by other online articles Search for citing articles ISI Web of Science (2) in: PubMed PubMed Citation Articles by Davey, P. A. Articles by Long, S. P. Agricola Articles by Davey, P. A. Articles by Long, S. P. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED ABSTRACT Averaged across many previous investigations, doubling the CO2 concentration ([CO2]) has frequently been reported to cause an instantaneous reduction of leaf dark respiration measured as CO2 efflux. No known mechanism accounts for this effect, and four recent studies have shown that the measurement of respiratory CO2 efflux is prone to experimental artifacts that could account for the reported response. Here, these artifacts are avoided by use of a high-resolution dual channel oxygen analyzer within an open gas exchange system to measure respiratory O2 uptake in normal air. Leaf O2 uptake was determined in response to instantaneous elevation of [CO2] in nine contrasting species and to long-term elevation in seven species from four field experiments. Over six hundred separate measurements of respiration failed to reveal any decrease in respiratory O2 uptake with an instantaneous increase in [CO2]. Respiration was found insensitive not only to doubling [CO2], but also to a 5-fold increase and to decrease to zero. Using a wide range of species and conditions, we confirm earlier reports that inhibition of respiration by instantaneous elevation of [CO2] is likely an experimental artifact. Instead of the expected decrease in respiration per unit leaf area in response to long-term growth in the field at elevated [CO2], there was a significant increase of 11% and 7% on an area and mass basis, respectively, averaged across all experiments. The findings suggest that leaf dark respiration will increase not decrease as atmospheric [CO2] rises. A quantitative analysis of prior studies weighted for replication and experimental variation concluded that a doubling of atmospheric CO2 concentration would decrease respiratory CO2 efflux by 18% in woody plants (Curtis and Wang, 1998 ). This decrease is considered the result of a direct instantaneous effect of increased CO2 concentration and a longer term indirect effect due to changes in leaf composition with long-term growth at elevated CO2 (for review, see Drake et al., 1999 ). In an analysis of 45 species, 36 showed an average 15% instantaneous reduction in net respiratory CO2 efflux per unit leaf area (Rd,CO2) on transfer to elevated [CO2] (for review, see Amthor, 1997 ). Terrestrial plant respiration releases 10 times more carbon per annum than fossil fuel combustion (Amthor, 1997 ). Therefore, a 15% to 20% decrease in foliar respiration might increase the potential of terrestrial vegetation to sequester carbon into biomass, providing a partial amelioration to the rate of increase of atmospheric [CO2] (for review, see Gonzalez-Meler and Siedow, 1999 ). However, there is considerable variation in the reported instantaneous effects of elevated [CO2] on Rd,CO2 with some studies reporting a large decrease and others reporting no change (Bunce, 2002 ; Bruhn et al., 2002 ). Furthermore, whereas some biochemical reactions within plant respiratory metabolic pathways are sensitive to [CO2], none of those identified exert sufficient metabolic control to account for the reported decreases in respiration (Gonzalez-Meler and Siedow, 1999 ). Four developments now suggest that the decreases observed when measuring Rd,CO2 from leaves in gas exchange systems incorporating infrared CO2 analyzers may be the result of artifacts in the measuring system. Amthor (2000 ) showed that very thorough sealing and use of an enlarged cuvette enclosing the leaf greatly reduced the apparent instantaneous effect of doubling [CO2] on Rd,CO2, although a small significant effect still remained. Jahnke (2001 ) systematically analyzed and eliminated potential artifacts within a gas exchange system and subsequently failed to detect any response in bean (Phaesoleus vulgaris) and poplar (Populus tremula) leaves when corrections were applied. Bouma et al. (1997 ) and Burton and Pregitzer (2002 ) similarly demonstrated that respiratory efflux of CO2 from roots was independent of measurement CO2 concentration. Assuming common dark respiratory metabolism between roots and leaves, this casts further doubt on the possibility of a direct response in the leaf. Avoiding the potential artifacts of CO2 exchange measurement by determining dark respiratory O2 uptake (Rd,O2), Amthor et al. (2001 ) showed no significant change in respiration of five Rumex crispus leaves in sharp contrast to earlier measurements of Rd,CO2 in the same species (Amthor et al., 1992 ). Measurement of O2 uptake has three important advantages over measurement of CO2 efflux for an instantaneous effect of change in [CO2] on respiration. First, the gas being measured is not the gas being altered in concentration, avoiding any need for instrument recalibration. Second, the concentration gradient of [O2] between the cuvette enclosing the leaf and the surrounding air is unaltered when [CO2] is changed. Finally, O2, unlike CO2, is not easily absorbed and adsorbed by surfaces in the gas exchange system. Yet, even when the effect of elevated [CO2] on O2 uptake has been examined, findings have been variable. In contrast to Amthor et al. (2001 ), Reuveni et al. (1993 ) measured a decrease in respiratory O2 uptake under elevated [CO2]. Controversy over the presence or absence of a direct effect of [CO2] on Rd,CO2 continues. Some recent studies of Rd,CO2, despite controlling for errors likely to arise from leaks, continue to report a significant instantaneous decrease in Rd,CO2 on elevation of [CO2]. For example, decreases have been reported of 19% for rice (Oryza sativa) canopies on elevation of [CO2] from 350 to 700 µmol mol-1 (Baker et al., 2000 ); 31% to 79% for Scots pine (Pinus sylvestris) needles also on elevation from 370 to 700 µmol mol-1 (Jach and Ceulemans, 2000 ); 16% to 40% for eight species of crop and crop weeds on increase from 60 to 1,000 µmol mol-1 (Bunce, 2001 ); an 11% average for leaves of 12 grass species on elevation from 360 to 1,300 µmol mol-1 (Tjoelker et al., 2001 ); 18% in the second leaf of soybean (Glycine max) on elevation from 350 to 1,400 µmol mol-1 (Bunce, 2002 ); and 4.5% decrease for seedling shoots of the tropical rainforest tree Tectona grandis on elevation from 370 to 600 µmol mol-1 (Holtum and Winter, 2003 ). Precautions taken in measurement were sufficient for Bunce (2002 ) to rebut the idea that the direct effect of elevated [CO2] on respiration was an artifact: “These precautions coupled with observed effects of dark [CO2] treatments on mass accumulation and translocation, suggest that the observed changes in respiration were real.” Although these decreases are smaller than those suggested by earlier studies, and some use larger increases in [CO2] to elicit an effect, they continue to suggest direct instantaneous inhibition of leaf dark respiration by increase in [CO2]. An added uncertainty is whether an instantaneous suppression of respiration by elevated [CO2] is limited to certain species or conditions (Drake et al., 1999 ). Although maintaining the same [CO2] outside the cuvette to that inside will minimize diffusive leaks, it will not eliminate the substantial problem of CO2 absorption/desorption from surfaces in the gas exchange system (Bloom et al., 1980 ; Long et al., 1996 ; Jahnke, 2001 ) or the mass flow of CO2 between the cuvette and the outside through the intercellular air space when only part of the leaf is enclosed (Jahnke and Krewitt, 2002 ). Thus, some of these further artifacts may still affect even these very careful studies. Clearly, we are some way from any consensus over the existence, or otherwise, of a direct effect of increase in [CO2] on leaf respiration in the dark. If the artifacts suggested by Amthor (2000 ), Jahnke (2001 ), and Jahnke and Krewitt (2002 ) apply to the effect of an instantaneous increase in [CO2] on Rd,CO2, then the same errors would at least in part apply when these systems are used to measure the effect of long-term acclimation to elevated [CO2] on Rd,CO2. Plants grown in elevated [CO2] in the long-term (1-10 years) typically show decreased nitrogen content and increased photosynthesis, growth, leaf mass per unit area, nonstructural carbohydrate concentrations (Drake et al., 1997 ), and large increases in leaf mitochondrial numbers (Robertson et al., 1995 ; Griffin et al., 2001 ; Tissue et al., 2002 ). With the exception of nitrogen content, these are all factors that might be expected to lead to increased leaf respiration. However, a quantitative survey using meta-analytical analysis concluded that long-term growth at elevated [CO2] decreased Rd,CO2 significantly and by 18% (Wang and Curtis, 2002 ). Recently, high-sensitivity dual lead-gold (fuel cell) detectors have been developed that allow measurement of small differences in [O2] (approximately 1-2 µmol mol-1) between the inlet and outlet air streams of a leaf cuvette in normal air (Willms et al., 1997 , 1999 ; Hunt, 2003 ). The objectives of this study were to use such a measurement system to answer two questions: Can the findings of Amthor et al. (2001 ) of no instantaneous effect of doubling of [CO2] on leaf Rd,O2 be extended to a wide range of species, growth conditions, and measurement [CO2]? Using a range of long-term experiments in which a wide range of plants have been grown under elevated [CO2] in the field (Table I) is Rd,O2 decreased or unaffected as a result of acclimation to elevated [CO2]? View this table: [in this window] [in a new window] Table I. Species, [CO2] treatments, plant and leaf developmental stage at the time of measurement Locations for material growing in the field and microclimate for controlled environment grown plants are given under “Materials and Methods”. References providing a description of each of the long-term field elevated [CO2] experiments used are included. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED RESULTS There was no significant instantaneous effect of [CO2] on leaf Rd,O2 in any of the nine species studied (Fig. 1; Tables II) and III). The results for the ANOVA were, for CO2, F3,332 = 0.032, P = 0.99; and for species x CO2, F24,332 = 0.528, P = 0.97. A complete lack of any effect of [CO2] was evident in all species not only on doubling [CO2], but when [CO2] was increased more than five times above the current ambient concentration or decreased to zero (Fig. 1; Tables II and III). The extremely high probabilities for accepting the null hypothesis (P > 0.97) eliminate any possibility of a Type II error, i.e. a difference undetected because of variability. This is also reflected in the combined regression of response relative to ambient [CO2] (Fig. 1), where the probability that the response is independent of [CO2] was P = 0.99 (F3,360 < 0.001). This lack of response of respiration to elevated [CO2] was independent of treatment method (F3,360 = 0.222, P = 0.88), developmental stage (F3,360 = 0.174, P = 0.92), beginning or end of night (F3,360 = 0.638, P = 0.59), and the [CO2] at which the plants had been grown (F3,100 = 0.080, P = 0.97). Although there was no interaction with [CO2], absolute rates of respiration generally decreased with time in the dark. Figure 1. The instantaneous effect of change in [CO2], from the current ambient concentration, on leaf respiratory O2 uptake in the dark. Respiration is plotted as the percentage of change (±1 SE) relative to the rate of O2 consumption at current ambient [CO2]. Each mean illustrated is based on measurements of 274 plants and nine species (Table I). Leaves were in darkness and their temperature was maintained at 25°C. Results for an analysis View larger version (13K): of variance (ANOVA) are given in the results text, and [in this window] respiration rates for individual species contributing to the [in a new window] illustrated means are shown in Tables II and III. View this table: [in this window] [in a new window] View this table: [in this window] [in a new window] Table II. The response of leaf respiration to changes in [CO2] for plants grown at current [CO2] only. The mean dark leaf O2 uptake on an area basis (micromoles per minute per second), and its SE (n = 6), are given for each of the four species at the beginning and end of the 10-h night. Table III. The response of leaf respiration to changes in (CO2) for species grown at long-term elevated (CO2) experiments The mean respiration rate and its SEM are given on an area basis and a dry weight basis. n indicates the number of plots or open-top chambers from which measurements of each treatment were made. Bold values show the critical longterm comparison of leaves grown and measured at their growth concentrations. An asterisk indicates that leaves grown and measured at elevated (CO2) have significantly different rates (P < 0.05) to those grown and measured at current ambient (CO2). Despite the lack of an instantaneous effect, long-term growth at elevated [CO2] did affect respiration. Leaves grown and measured at elevated [CO2] had the same or a significantly higher Rd,O2 than those grown and measured at current ambient [CO2] (Table III). The largest increases in respiration were for Quercus geminata (23%) and soybean (22%) on a leaf area basis. By contrast, there was no increase in Acer saccharum, Betula papyrifera, and Populus tremuloides (Table IIIa). When all values from the long-term experiments were combined, there was a significant increase in Rd,O2 of 11% and 7% on a leaf area and mass basis, respectively, and relative to controls (Fig. 2). Figure 2. The effect of long-term (lifetime or >3 years) growth at elevated [CO2] in the field on leaf respiration. The data points show mean leaf respiratory O2 uptake (±90% confidence intervals) on an area and mass basis for View larger version (8K): leaves grown at elevated [CO2] relative to controls grown at [in this window] current ambient [CO2]. The points are means for seven [in a new window] species (Table I) weighted for replication and variance using meta-analytical statistical methods. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED DISCUSSION Six hundred separate measurements of different leaves, encompassing four long-term field CO2 enrichment experiments revealed no evidence of any decrease in Rd,O2 in response to instantaneous elevation of [CO2]. This was consistent for seven tree and two crop species, the majority sampled from long-term elevated [CO2] experiments where plants have developed throughout their life or for several seasons under elevated [CO2]. This lack of response was irrespective of developmental stage, growth conditions, and growth [CO2]. Bunce (2001 ) reported that the sensitivity of Rd,CO2 to an instantaneous elevation of [CO2] increased with duration of the dark period. However, no effect on Rd,O2 was evident here (Tables II and III), regardless of whether measurements were made at the start or end of the natural dark period. Jahnke (2001 ) similarly found no effect of [CO2] on Rd,CO2 even after 72 h of darkness. Tjoelker et al. (2001 ) noted that the instantaneous decrease in Rd,CO2 with elevated [CO2], although minor (1.8%) in response to increase from 360 to 700 µmol mol-1, was far more substantial (11%) on increase from 360 to 1,300 µmol mol-1. However, a 5-fold increase above current ambient [CO2] or removal of CO2 from the atmosphere around the leaf did not alter Rd,O2 (Fig. 1). These findings support the conclusions of Amthor et al. (2001 ) and of Jahnke (2001 ) and extend them to more species and to major long-term [CO2] field manipulation studies. Although the results show no effect of instantaneous elevation of [CO2] on Rd,O2, this cannot rule out some effect on apparent Rd,CO2 due to non-respiratory processes, for example, CO2 uptake into carboxylic acids. However, in one of the very few studies to measure CO2 and O2 exchange at different [CO2], Amthor et al. (2001 ) found no effect of elevated [CO2] on respiratory quotient. It has been suggested that the variability of response to instantaneous increase in [CO2] apparent in the literature may be the result of species differences (Drake et al., 1999 ; Baker et al., 2000 ; Hamilton et al., 2001 ). Without definitive testing, it is impossible to be certain on this point. However, this study contains a diverse range of species and functional types (including C3 and C4 photosynthetic types, herbaceous and woody forms, two major crops, and seven forest dominant gymnosperm and angiosperm species) that individual exceptions exist now seem somewhat improbable. Moreover, we deliberately selected species previously reported to show some of the largest inhibitions of respiration by elevated CO2 with previous methods and technology: soybean 40% (Bunce, 1995 ), Pinus taeda 14% (Teskey, 1995 ), Quercus rubra 5.6% (Amthor, 2000 ), and maize (Zea mays) 46% (Cornic and Jarvis, 1972 ). Any errors associated with measuring the instantaneous response of respiratory CO2 efflux to elevated [CO2] may be equally applicable to measurement of the long-term effects of growth at elevated [CO2] on Rd,CO2. The significant 7% increase in Rd,O2 on a mass basis, averaged across seven species from four long-term experiments (Fig. 2), contrasts sharply with the average 18% decrease reported by Wang and Curtis (2002 ) in their meta-analysis of Rd,CO2 on a mass basis measured in long-term experiments. The significant increase on a mass, as well as an area basis, suggests that increased respiration per unit leaf area is not just the result of the increased mass per unit area that is commonly observed in leaves grown at elevated [CO2] (Drake et al., 1999 ). However, as noted above, this mean increase masks marked differences between species in this long-term response (Table III). To our knowledge, this is the first study to use this alternative method to investigate the effects of long-term growth at elevated [CO2] on respiration. The increase in respiration shown here is consistent with the widely reported increase in leaf-soluble carbohydrates with long-term growth at elevated [CO2] (Moore et al., 1999 ). Azcon-Bieto and Osmond (1983 ) showed that increase in leaf nonstructural carbohydrate content produced by manipulation of photosynthesis during the photoperiod increased subsequent dark respiration rates. Therefore, a similar response might be expected when elevated [CO2] increases photosynthesis and leaf nonstructural carbohydrate content during the photoperiod. In conclusion, instantaneous elevation of [CO2] had no effect on leaf O2 uptake in these nine species. This further confirms the suggestion that reports of respiratory inhibition, based on measurements of CO2 uptake, are the result of experimental artifacts and not the result of any sensitivity of plant respiration to the [CO2] at the time of measurement. A 15% to 20% reduction of terrestrial plant respiration with a doubling of atmospheric [CO2] concentration would represent some amelioration, at least temporarily, in the rate of rise in the global atmospheric [CO2] (for review, see Gonzalez-Meler and Siedow, 1999 ). It can no longer be assumed that the direct effect of elevated [CO2] on plant respiration will reduce future ecosystem CO2 efflux. On the basis of the long-term field manipulations of [CO2] studied here, a small increase in respiration per unit leaf area should be expected—this will amplify the effect of any increase in leaf area due to growth at elevated [CO2]. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED MATERIALS AND METHODS Growth Conditions The species, developmental stage, growth conditions, and experimental site are listed in Table I. Seven out of nine species were grown at current ambient (368 µmol mol-1) and at an elevated [CO2] (550 or 700 µmol mol-1). Two of the species were measured at more than one developmental stage; soybean (Glycine max) at vegetative and podfill stages, and Pinus taeda at juvenile and mature stages. Soybean was grown in controlled environment- and open-field conditions (FACE). Oxygen Analysis System Leaf O2 uptake rate was determined in an open gas exchange system incorporating a dual-cell oxygen analyzer (S-3A/DOX; AEI Technologies, Pittsburgh) described by Willms et al. (1997 ), and incorporating a stainless steel and glass leaf gas exchange cuvette (MPH-1000; Campbell Scientific, Logan, UT). The analyzer incorporates a novel differential oxygen sensor obtained by modifying two electrochemical fuel cells (model KE-25; Figaro USA, Wilmette, IL), placing one in a reference gas stream and the other in an analytical gas stream, and then ensuring that the two gas sensors were exposed to an identical temperature and pressure (Willms et al., 1997 ). The O2 fuel cells contain a Teflon membrane through which O2 can diffuse, and a weak acid electrolyte bathes a lead anode and a gold cathode electrode. In the environment of the sensor, the O2 oxidizes the lead, forming PbO, which dissolves into the solution and generates a small current. The difference in output of the two sensors is directly proportional to the [O2] between the air streams. The selective diffusive properties of Teflon and the specific electrochemistry ensure that the differential sensor responds directly only to O2 in air. However, changes in partial pressures of CO2 and water vapor due to respiration and transpiration will affect that of O2 causing an indirect effect (Willms et al., 1999 ). For this reason, these gases are removed from the air stream before the O2 sensors (Hunt, 2003 ). The analyzer was secured to a vibrationless table top and additional layers of insulation were wound around the O2 measurement cells and instrument casing to further improve resolution. The differential resolution of the modified dual-cell O2 analyzer was 1 to 2 µmol mol-1 against a background [O2] of 210 mmol mol-1. The cuvette was large enough to enclose the entire leaf, leaflet, or cohort of needles; typically, 50 to 120 cm2 of leaf were enclosed. This avoided errors associated with gas flow between the cuvette and surrounding air via the intercellular air space that Jahnke and Krewitt (2002 ) have shown to result from the use of cuvettes that only enclose part of the leaf. The cuvette (MPH-1000; Campbell Scientific) has been described in detail previously (Bingham et al., 1980 ), adapted from the earlier design of Bingham and Coyne (1977 ). Typical boundary layer conductances to water vapor are 2.5 to 4 mol m-2 s-1 (calculated from Bingham and Coyne, 1977 ; Bingham et al., 1980 ). A gas exchange system (LI6400; LI-COR, Lincoln, NE) was used to provide a controlled [CO2] to the cuvette, and air flow rate was measured by a pair of flow meters (F350; AEI Technologies), calibrated as described previously (Long et al., 1996 ). The gas flow across the leaf was maintained at 2.5 cm3 s-1. This flow rate and the enclosed leaf area resulted in a [O2] of 40 to 80 µmol mol-1. The entry air streams to the oxygen analyzer were scrubbed of water and CO2 by magnesium perchlorate and soda lime. Column sizes were calculated to remove all CO2 and to be sufficient to dry the air stream to the equilibrium vapor pressure of magnesium perchlorate (approximately 0.1 Pa). The efficacy of these columns was tested by switching dry air (<0.01 kPa water vapor) with a humidified air stream (2.0 kPa water vapor) at the cuvette inlet; no change in [O2] from 0 µmol mol-1 was recorded. Similar tests showed that the soda lime column was effective in preventing interference from differences in [CO2] between the cuvette and reference air streams. The response time of the gas exchange system was assessed using two cylinders of compressed air found to differ in [O2] by 80 µmol mol-1. Air from one of these cylinders was fed to the empty cuvette inlet and also the reference air stream until a constant [O2] of 0 µmol mol-1 was obtained at the analyzer. Air from the second cylinder was then switched in at the cuvette inlet and the time taken to obtain a constant [O2] of 80 µmol mol-1 at the analyzer was recorded. The change was 98% complete within 10 min and complete within the resolution of the analyzer at 15 min. Therefore, it was assumed that 20 min was adequate to record any response of respiratory O2 efflux to an instantaneous change in [CO2]. Respiration Measurements Respiration was measured in the last 2.5 h of the night in all species and additional measurements were made in the first 2.5 h of the night for three species (Table I). Measurements were made at each location listed in Table I. However, because of its size and stability, the gas exchange system was housed in a field laboratory at each site and leaves were detached shortly before measurement. Petioles were cut under water and their cut ends were kept immersed in water until measurements were complete. The leaves remained turgid throughout. For pot-grown soybean (Table I), parallel measurements on attached and detached leaves were made; no significant differences in respiration or its response to [CO2] were found. Leaves were equilibrated in the dark to 25°C before measurement. For a single replicate measurement, respiration was measured at a [CO2] of 2,000 µmol mol-1, and then again at 700, 550, 360, and 0 µmol mol-1 CO2. To avoid systematic error, each alternate replicate measurement was started at 0 µmol mol-1 CO2 and was then switched to 368, 700, 550, and 2,000 µmol mol-1 CO2. Species from the FACE experiments at Duke Forest, Rhinelander, and the University of Illinois, and the open-top chamber experiment at Merritt Island were measured at 550 µmol mol-1 [CO2] rather than at 700 µmol mol-1 [CO2] because this was the growth concentration (Table I). After each step-wise change in [CO2], a 20min waiting period ensured adequate time for any instantaneous response in Rd,O2. Each set of measurements took 80 min to complete, and leaves were maintained in the dark at a leaf temperature of 25°C for the entire measurement period. After each set of measurements, leaf area was measured by digital imaging and the leaves were then dried to a constant weight at 80°C. Leaf O2 evolution rates were calculated on a leaf area and a dry mass basis by adapting the CO2 uptake equations of von Caemmerer and Farquhar (1981 ) for open system gas exchange measurements. Statistical Analysis In the four long-term field experiments, replicate number was determined by the number of treatment plots (Table III). This avoided pseudoreplication. A minimum of three repeat measurements were taken in each treatment plot of each experiment, and their pooled values provided the individual sample measure. Only these data were used to assess the effect of longterm growth at elevated [CO2] on respiratory O2 uptake. Outside of these long-term elevated [CO2] studies, measurements were made from six randomly selected plants each of P. taeda seedlings, Q. rubra mature trees, and maize flowering plants growing under current ambient [CO2] in the open. Soybean was grown in two separate controlled environment studies, and here, replicates represent the number of separate chambers used (Table I). ANOVA was used to test the effect of immediate changes of [CO2] on leaf respiration rate for all nine species combined (SYSTAT, Evanston, IL). This tested the effect of species, growth conditions, developmental stage, time of measurement, and growth [CO2] on the instantaneous response to elevated [CO2]. In addition, a regression analysis of response relative to ambient [CO2] was tested (SYSTAT). Because of the very different experimental designs, durations, and treatment procedures of the experiments in which plants were grown at elevated [CO2], and the possibility that species may respond differently to long-term growth in elevated [CO2], the pooled effect of all field studies was assessed by meta-analytical statistical techniques (Hedges et al., 1999 ; Ainsworth et al., 2002 ; Morgan et al., 2003 ) using a software package (MetaWin, Rosenberg et al., 2000). This allows determination of the consensus of studies differing in design. The natural log of the response ratio (respiratory O2 uptake for leaves grown at elevated [CO2]/respiratory O2 uptake for leaves grown at the current ambient [CO2]) was used as the metric for analysis (Hedges et al., 1999 ). A mixed-model analysis was used based on the assumption of random variation in effect sizes between individual measurements. A weighted parametric analysis was used (Gurevitch and Hedges, 1999 ), and each individual observation was weighted by the reciprocal of the mixedmodel variance (Hedges et al., 1999 ). If the response ratio was not significantly different from unity, growth at elevated [CO2] failed to change leaf O2 uptake from that at current ambient [CO2]. However, if significantly different from 1, growth at elevated [CO2] on average altered O2 uptake across the studies. ACKNOWLEDGMENTS We thank Henry Ginsberg and Joseph Veltre of AEI Technologies, and Nicholas Dowling of Qubit Systems for technical help and modifications to the O2 analysis system; Jeff Amthor, Miquel Gonzalez-Meler, Andrew Leakey, and Shawna Naidu for comments on draft versions of the manuscript; Carlos Pimentel and Kate George for the supply of plant material; and Lisa Ainsworth for advice on the meta-analysis. Received July 20, 2003; returned for revision August 20, 2003; accepted September 25, 2003. FOOTNOTES 1 This work was supported by the U.S. Department of Energy’s Office of Science (to B.E.R.), as were the long-term elevated [CO2] experiments in Florida, North Carolina, and Wisconsin. 2 Present address: Department of Biological Sciences, University of Essex, Colchester, CO4 3SQ, UK. 3 Present address: Dipartimento di Scienze dell’ Ambiente Forestale e delle Sue Risorse, Università della Tuscia, Via S. Camillo de Lellis, 01100 Viterbo, Italy. Article, publication date, and citation information www.plantphysiol.org/cgi/doi/10.1104/pp.103.030569. • can be found at Corresponding author; e-mail stevel@life.uiuc.edu; fax 217-244-7563. TOP ABSTRACT RESULTS DISCUSSION MATERIALS AND METHODS LITERATURE CITED LITERATURE CITED Ainsworth EA, Davey PA, Bernacchi CJ, Dermody OC, Heaton EA, Moore DJ, Morgan PB, Naidu SL, Yoo Ra HS, Zhu XG et al. (2002) A meta-analysis of elevated [CO2] effects on soybean (Glycine max) physiology, growth and yield. Glob Change Biol 8: 695709[CrossRef][ISI] Ainsworth EA, Tranel PJ, Drake BG, Long SP (2003) The clonal structure of Quercus geminata revealed by conserved microsatellite loci. Mol Ecol 12: 527532[CrossRef][ISI][Medline] Amthor JS (1997) Plant respiratory responses to elevated CO2 partial pressure. In LH Allen, MB Kirkham, DM Olszyk, CE Whitman, eds, Advances in Carbon Dioxide Effects Research. American Society of Agronomy Special Publication, Madison, WI, pp 35-77 Amthor JS (2000) Direct effect of elevated CO2 on nocturnal in situ leaf respiration in nine temperate deciduous tree species is small. Tree Physiol 20: 139-144[ISI][Medline] Amthor JS, Koch GW, Bloom AJ (1992) CO2 inhibits respiration in leaves of Rumex crispus L. Plant Physiol 98: 757-760[ISI] Amthor JS, Koch GW, Willms JR, Layzell DB (2001) Leaf O2 uptake in the dark is independent of coincident CO2 partial pressure. J Exp Bot 52: 2235-2238[Abstract/Free Full Text] Azcon-Bieto L, Osmond CB (1983) Relationship between photosynthesis and respiration: the effect of carbohydrate status on the rate of CO2 production by respiration in darkened and illuminated wheat leaves. Plant Physiol 71: 574-581[ISI] Baker JT, Allen LH, Boote KJ, Pickering NB (2000) Direct effects of atmospheric carbon dioxide concentration on whole canopy dark respiration of rice. Glob Change Biol 6: 275286[CrossRef][ISI] Bingham GE, Coyne PI (1977) A portable, temperature-controlled, steady-state porometer for field measurements of transpiration and photosynthesis. Photosynthetica 11: 148-160[ISI] Bingham GE, Coyne PI, Kennedy RB, Jackson WL (1980) Design and fabrication of a portable minicuvette system for measuring leaf photosynthesis and stomatal conductance under controlled conditions. UCRL-52895. Lawrence Livermore National Laboratory, Livermore, CA Bloom AJ, Mooney HA, Bjorkman O, Berry JA (1980) Materials and methods for carbon dioxide and water exchange analysis. Plant Cell Environ 3: 371-376[ISI] Bouma TJ, Nielsen KL, Eissenstat DM, Lynch JP (1997) Soil CO2 concentration does not affect growth or root respiration in bean or citrus. Plant Cell Environ 20: 14951505[CrossRef][ISI] Bruhn D, Mikkelsen TN, Atkin OK (2002) Does the direct effect of atmospheric CO2 concentration on leaf respiration vary with temperature? Responses in two species of Plantago that differ in relative growth rate. Physiol Plant 114: 57-64[CrossRef][ISI][Medline] Bunce JA (1995) The effect of carbon dioxide concentration on respiration of growing and mature soybean leaves. Plant Cell Environ 18: 575-581[ISI] Bunce JA (2001) Effects of prolonged darkness on the sensitivity of leaf respiration to carbon dioxide concentration in C3 and C4 species. Ann Bot 87: 463-468[Abstract/Free Full Text] Bunce JA (2002) Carbon dioxide concentration at night affects translocation from soybean leaves. Ann Bot 90: 399-403[Abstract/Free Full Text] Burton AJ, Pregitzer KS (2002) Measurement carbon dioxide concentration does not affect root respiration of nine tree species in the field. Tree Physiol 22: 67-72[ISI][Medline] Cornic G, Jarvis PG (1972) Effects of oxygen on CO2 exchange and stomatal resistance in Sitka spruce and maize at low irradiances. Photosynthetica 6: 225-239[ISI] Curtis PS, Wang X (1998) A meta-analysis of elevated CO2 effects on woody plant mass, form and physiology. Oecologia 113: 299-313[CrossRef][ISI] DeLucia EH, Hamilton JG, Naidu SL, Thomas RB, Andrews JA, Finzi A, Lavine M, Matamala R, Mohan JE, Hendrey GR et al. (1999) Net primary production of a forest ecosystem with experimental CO2 enrichment. Science 284: 1177-1179[Abstract/Free Full Text] Dickson RE, Lewin KF, Isebrands JG, Coleman MD, Heilman WE, Riemenschneider DE, Sober J, Host GE, Hendrey GR, Pregitzer KS et al. (2001) Forest atmosphere carbon transfer storage-II (FACTS II): the aspen free-air CO2 and O3 enrichment (FACE) project in an overview. General Tech. Rep. NC-214. U.S. Department of Agriculture Forest Service North Central Experiment Station, Rhinelander, WI. Drake BG, Azcon-Bieto J, Berry J, Bunce J, Dijkstra J, Farrar J, Gifford RM, GonzàlezMeler MA, Koch G, Lambers H et al. (1999) Does elevated atmospheric CO2 concentration inhibit mitochondrial respiration in green plants? Plant Cell Environ 22: 649-657[CrossRef][ISI] Drake BG, Gonzalez-Meler M, Long SP (1997) More efficient plants: a consequence of rising atmospheric CO2. Annu Rev Plant Physiol Plant Mol Biol 48: 607-637 Gonzalez-Meler MA, Siedow JN (1999) Direct inhibition of mitochondrial respiratory enzymes by elevated CO2: Does it matter at the tissue or whole-plant level? Tree Physiol 19: 253259[ISI][Medline] Griffin KL, Anderson OR, Gastrich MD, Lewis JD, Lin GH, Schuster W, Seemann JR, Tissue DT, Turnbull MH, Whitehead D (2001) Plant growth in elevated CO2 alters mitochondrial number and chloroplast fine structure. Proc Natl Acad Sci USA 98: 24732478[Abstract/Free Full Text] Gurevitch J, Hedges LV (1999) Statistical issues in ecological meta-analyses (meta-analysis in ecology). Ecology 80: 1142-1150[ISI] Hamilton JG, Thomas RB, Delucia EH (2001) Direct and indirect effects of elevated CO2 on leaf respiration in a forest ecosystem. Plant Cell Environ 24: 975-982[CrossRef][ISI] Hedges LV, Gurevitch J, Curtis PS (1999) The meta-analysis of response ratios in experimental ecology. Ecology 80: 1150-1156[ISI] Holtum JAM, Winter K (2003) Photosynthetic CO2 uptake in seedlings of two tropical tree species exposed to oscillating elevated concentrations of CO2. Planta 218: 152158[CrossRef][ISI][Medline] Hunt S (2003) Measurements of photosynthesis and respiration in plants. Physiol Plant 117: 314325[CrossRef][ISI][Medline] Hymus GJ, Snead TG, Johnson DP, Hungate BA, Drake BG (2002) Acclimation of photosynthesis and respiration to elevated atmospheric CO2 in two Scrub Oaks. Glob Change Biol 8: 317-328[CrossRef][ISI] Jach ME, Ceulemans R (2000) Short-versus long-term effects of elevated CO2 on night-time respiration of needles of Scots pine (Pinus sylvestris L.). Photosynthetica 38: 5767[CrossRef][ISI] Jahnke S (2001) Atmospheric CO2 concentration does not directly affect leaf respiration in bean or poplar. Plant Cell Environ 24: 1139-1151[CrossRef][ISI] Jahnke S, Krewitt M (2002) Atmospheric CO2 concentration may directly affect leaf respiration measurement in tobacco, but not respiration itself. Plant Cell Environ 25: 641-651[CrossRef][ISI] Long SP, Farage PK, Garcia RL (1996) Measurement of leaf and canopy photosynthetic CO2 exchange in the field. J Exp Bot 47: 1629-1642[Abstract] Moore BD, Cheng SH, Sims D, Seemann JR (1999) The biochemical and molecular basis for photosynthetic acclimation to elevated atmospheric CO2. Plant Cell Environ 22: 567582[CrossRef][ISI] Morgan PB, Ainsworth EA, Long SP (2003) How does elevated ozone impact soybean? A meta-analysis of photosynthesis, growth and yield. Plant Cell Environ 26: 13171328[CrossRef][ISI] Reuveni J, Gale J, Mayer AM (1993) Reduction of respiration by high ambient CO2 and the resulting error in measurement of respiration made with O2 electrodes. Ann Bot 72: 129131[Abstract/Free Full Text] Robertson EJ, Williams M, Harwood JL, Lindsay JG, Leaver CJ, Leech RM (1995) Mitochondria increase 3-fold and mitochondrial proteins and lipid change dramatically in postmeristematic cells in young wheat leaves grown in elevated CO2. Plant Physiol 108: 469474[Abstract/Free Full Text] Rogers A, Allen DJ, Davey PA, Morgan PB, Ainsworth EA, Bernacchi CJ, Cornic G, Dermody O, Heaton EA, Mahoney J et al. (2004) Leaf photosynthesis and carbohydrate dynamics of soybeans grown throughout their life-cycle under free-air carbon dioxide enrichment. Plant Cell Environ 27 (in press) Rosenburg MS, Adams DC, Gurevitch J (2000) MetaWin: Statistical Software for MetaAnalysis, Version 2.0. Sinauer Associates, Sunderland, MA Teskey RO (1995) A field study of the effects of elevated CO2 on carbon assimilation, stomatal conductance and leaf and branch growth of Pinus taeda trees. Plant Cell Environ 18: 565-573[ISI] Tissue DT, Lewis JD, Wullschleger SD, Amthor JS, Griffin KL, Anderson R (2002) Leaf respiration at different canopy positions in sweetgum (Liquidambar styraciflua) grown in ambient and elevated concentrations of carbon dioxide in the field. Tree Physiol 22: 11571166[ISI][Medline] Tjoelker MG, Oleksyn J, Lee TD, Reich PB (2001) Direct inhibition of leaf dark respiration by elevated CO2 is minor in 12 grassland species. New Phytol 150: 419-424[CrossRef][ISI] von Caemmerer S, Farquhar GD (1981) Some relationships between the biochemistry of photosynthesis and the gas exchange of leaves. Planta 153: 376-387[ISI] Wang XZ, Curtis P (2002) A meta-analytical test of elevated CO2 effects on plant respiration. Plant Ecol 161: 251-261[CrossRef][ISI] Willms JR, Dowling AN, Dong ZM, Hunt S, Shelp BJ, Layzell DB (1997) The simultaneous measurement of low rates of CO2 and O2 exchange in biological systems. Anal Biochem 254: 272-282[CrossRef][ISI][Medline] Willms JR, Salon C, Layzell DB (1999) Evidence for light-stimulated fatty acid synthesis in soybean fruit. Plant Physiol 120: 1117-1127[Abstract/Free Full Text] This article has been cited by other articles: M. A. GONZALEZ-MELER, L. TANEVA, and R. J. TRUEMAN Plant Respiration and Elevated Atmospheric CO2 Concentration: Cellular Responses and Global Significance Ann. Bot., November 1, 2004; 94(5): 647 656. [Abstract] [Full Text] [PDF] This Article Abstract Full Text (PDF) All Versions of this 134/1/520 most pp.103.030569v1 Article: recent Alert me when this article is cited Alert me if a correction is posted Services Similar articles in this journal Similar articles in ISI Web of Science Similar articles in PubMed Alert me to new issues of the journal Download to citation manager Search for citing articles ISI Web of Science (2) PubMed PubMed Citation Articles by Davey, P. A. Articles by Long, S. P. Agricola Articles by Davey, P. A. Articles by Long, S. P. in: HOME HELP FEEDBACK ASPB JOURNALS SUBSCRIPTIONS ARCHIVE PLANT PHYSIOLOGY SEARCH TABLE OF CONTENTS THE PLANT CELL Copyright © 2004 by the American Society of Plant Biologists