Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Precipitation Regime Shift Enhanced the Rain Pulse Effect on Soil Respiration in a Semi-Arid Steppe

  • Liming Yan,

    Affiliations State Key Laboratory of Vegetation and Environmental Change, Institute of Botany, Chinese Academy of Sciences, Beijing, China, School of Life Sciences, Fudan University, Shanghai, China

  • Shiping Chen ,

    spchen@ibcas.ac.cn

    Affiliation State Key Laboratory of Vegetation and Environmental Change, Institute of Botany, Chinese Academy of Sciences, Beijing, China

  • Jianyang Xia,

    Affiliation Department of Microbiology and Botany, University of Oklahoma, Norman, Oklahoma, United States of America

  • Yiqi Luo

    Affiliation Department of Microbiology and Botany, University of Oklahoma, Norman, Oklahoma, United States of America

Abstract

The effect of resource pulses, such as rainfall events, on soil respiration plays an important role in controlling grassland carbon balance, but how shifts in long-term precipitation regime regulate rain pulse effect on soil respiration is still unclear. We first quantified the influence of rainfall event on soil respiration based on a two-year (2006 and 2009) continuously measured soil respiration data set in a temperate steppe in northern China. In 2006 and 2009, soil carbon release induced by rainfall events contributed about 44.5% (83.3 g C m−2) and 39.6% (61.7 g C m−2) to the growing-season total soil respiration, respectively. The pulse effect of rainfall event on soil respiration can be accurately predicted by a water status index (WSI), which is the product of rainfall event size and the ratio between antecedent soil temperature to moisture at the depth of 10 cm (r2 = 0.92, P<0.001) through the growing season. It indicates the pulse effect can be enhanced by not only larger individual rainfall event, but also higher soil temperature/moisture ratio which is usually associated with longer dry spells. We then analyzed a long-term (1953–2009) precipitation record in the experimental area. We found both the extreme heavy rainfall events (>40 mm per event) and the long dry-spells (>5 days) during the growing seasons increased from 1953–2009. It suggests the shift in precipitation regime has increased the contribution of rain pulse effect to growing-season total soil respiration in this region. These findings highlight the importance of incorporating precipitation regime shift and its impacts on the rain pulse effect into the future predictions of grassland carbon cycle under climate change.

Introduction

Global precipitation has been predicted to change with increasing intra-annual variability and more frequent extreme rainfall events [1], [2]. Such shifts in precipitation regime could have profound impacts on belowground carbon (C) release, especially in the arid and semiarid ecosystems [3][6]. It has been widely reported that soil respiration in grassland ecosystems can increase significantly and immediately after a rainfall event, followed by a decrease with declining soil moisture [4], [7]. This pulse effect of rainfall event on soil respiration has been suggested as an important contributor to ecosystem C release in grassland ecosystems [3], [8], [9]. Therefore, a better understanding of how the changing precipitation regime will affect rain pulse effect is important for predicting future grassland C feedbacks to climate change.

Previous studies have summarized that variations in soil respiration are determined by several factors, including soil temperature, soil water availability and carbon substrate supply [10][12]. Yet it is not clear which factors control the pulse effect of a single rainfall event on soil respiration. There have been both laboratory [13][15] and field [7], [16], [17] attempts trying to identify the regulatory mechanisms of rain pulses on soil respiration. The magnitude of soil respiration response (e.g., soil carbon release) is positively correlated with rainfall event size [4], [18], [19]. Other studies (e.g. [20]) have observed that the antecedent soil water condition is also an important influencing factor. The pulse effect is intensified by the dry condition of the antecedent soil [20], and the effect is less obvious if the soil is wet before the rainfall event [21][23]. Soil temperature has been widely reported to regulate soil microbial activities and thus soil heterotrophic respiration in various ecosystems [22], [24]. Other conditions, such as plant activity [25], [26] and soil organic matter content [20], [27], are also believed to influence the effects of rainfall event on soil respiration. Therefore, the predictability of rain pulse effect on soil respiration is still low and no effective approach or indicator has been developed so far.

In semi-arid ecosystems, water availability is the dominant factor regulating soil respiration [28][30]. Water availability also moderates the effects of the other factors on soil respiration such as temperature and substrate supply [30]. Water availability and its intra- and inter-annual variations are directly linked to both the intensity and the frequency of precipitation. Although IPCC [2] has reported a trend of significant change in both total amount and temporal patterns of precipitation, only a few studies have analyzed the general changes of precipitation regimes especially in semi-arid temperate grasslands [31]. Shifts in precipitation regime will alter not only the size of individual rainfall event, but also the length of the dry-spell duration and thus the antecedent soil water condition. Previous laboratory experiments have shown that both of these two factors can significantly influence the soil CO2 release [32], [33]. A growing body of works using field experiments have demonstrated that shifts in precipitation regime alone, even when the total rainfall amount does not change, can have large impacts on grassland soil CO2 release [29], [34], [35] and its pulse responses to rainfall events [4], [17]. However, to our knowledge, few studies have observed and quantified the dynamics of soil respiration after rainfall events [36], [37]. Therefore, it is necessary to incorporate the changes in precipitation regime when evaluating the effect of rainfall event on C cycling in natural grassland ecosystems.

The area of grasslands contributed to about 40% of the total territory of China [38], and the temperate steppe in China is the third largest grassland area in the world [39]. In this study, we analyzed both a long-term (1953–2009) rainfall data set and two years (2006 and 2009) of continuously measured soil respiration data in a temperate steppe in northern China. We attempt to address the following questions: (1) What are the controlling factors of the rainfall event effect on soil respiration through the growing season in this ecosystem? (2) Has precipitation regime changed over the last ∼60 yr in this semi-arid steppe? (3) How shifting precipitation regime would influence soil respiration responses to pulses in water availability?

Materials and Methods

2.1. Site description

Our field site (42°27′N, 116° 41′E), located Duolun County in the northeastern Inner Mongolia, China, belongs to Duolun Restoration Ecology Experimentation and Demonstration Station (DREEDS), Institute of Botany. No specific permissions were required for scientific researches. The site has been fenced in to exclude grazing since 2001, and no other management like mowing or fertilizing was applied. The mean altitude is about 1430 m above sea level. The vegetation was dominated by C3 grasses (e.g. Stipa krylovii, Agropyron cristatum, Leymus chinensis) and a semi-shrub species (Artemisia frigida). The soil in our study site is classified as Calcic-orthic Aridisol. The mean annual air temperature is 3.3°C and the mean annual precipitation is 377 mm, 95% (i.e. 358 mm) of which occurs during the growing season from April to October. Average air temperature was 3.13°C in 2006 and 3.09°C in 2009. Annual precipitation was 425 mm in 2006 and 185 mm in 2009, respectively.

2.2 Variable measurement

The growing season in this ecosystem usually starts from late April and ends in early October. Five PVC collars (20.3 cm in diameter and 8 cm in height) were inserted into the soil to a depth of 3 cm. The PVC collars were randomly distributed in the study site with 5 m–30 m distance between any two of them in late April, 2006, and covered an area of about 300 m2. The site is flat so topography had little impact on the difference in measured soil respiration among replicates. Soil respiration rates were continuously measured during the growing season (from May to October) of 2006 and 2009 on a half-hour basis using an infra-red gas analyzer (LI-840, Li-Cor Inc., Lincoln, NE, USA) that was connected to five automatic measurement chambers. All living plants inside the chambers were clipped by hand weekly to exclude aboveground plant respiration. The clipped aboveground plants were left in the chambers to include CO2 efflux from the litter decomposition. The CO2 concentrations in the chambers were recorded in CR1000 data logger (Campbell Scientific Inc., CSI, Utah, USA), and were processed by LoggerNet 3.1 (CSI, USA). Each measurement took 120 s. The soil respiration rates were determined from the time series of soil CO2 efflux concentrations. Since the measuring system worked well during 2006 and 2009, only small data gaps (less than 2 h) existed in two years data sets and were filled by linear interpolation method.

Soil temperatures (°C) at a depth of 10 cm and volumetric soil moisture (%) at a depth of 0–10 cm were measured using 107 soil temperature probes (CSI, USA) and CS616 soil water probes (CSI, USA), respectively. Three combinations of soil temperature and water probes were placed close (about 1 m) to the soil respiration chambers. The mean values of half-hour soil temperature and moisture were recorded in a CR1000 data logger (CSI, USA) simultaneously.

The precipitation data of the past 57 years were provided by local meteorological stations in Duolun County. The precipitation data of the growing season of 2006 and 2009 were measured by a tipping bucket rain gauge (TE525MM, CSI) at the study site.

2.3. Statistical analyses

Considering the significant impact of soil temperature on the diurnal variation of soil respiration, we analyzed precipitation and soil respiration data on a daily time step. There was a rainfall event when precipitation was recorded, and a day without any precipitation was defined as a dry day. The rainfall event size was defined as the total amount of rainfall during a rainfall event. Note that one day may have more than one rainfall event and, at the other extreme, one rainfall event may last for a few days. An uninterrupted sequence of dry days preceded and succeeded by at least one rainfall event is referred to as a dry-spell event [40], [41]. We added up the total precipitation amount and calculated the frequencies of different dry-spell durations (days) in growing seasons of the past 57 years (1953–2009). The rainfall and dry-spell events were then classified into different categories based on their size (rainfall event size: 0–2, 2–5, 5–10, 10–15, 15–20, 20–30, 30–40, 40–50, or >50 mm) and duration (dry-spell duration: 0–5, 5–10, 10–20, 20–30, or >30 days).

Since the size of individual rainfall events varies greatly, the rain pulse effect on soil respiration can last from several hours to a few weeks. Thus, if the rain pulse response lasted for at least three days, we would use daily soil respiration data with the following function to estimate the effect of a rainfall event on soil respiration [7]:(1)where yt is daily soil respiration rate after the rainfall event, y0 is initial daily soil respiration before the rainfall event, t is the time (continuous but not a discrete daily time-step) after the rainfall event, and a and b are coefficients. However, if the rain pulse response lasted less than three days, we would compare the difference between post-rainfall soil respiration and the control at half-hourly time scale. The sum of all the differences over one day greater than 0±0.01 (g C m−2 day−1) was defined as rainfall event. The calculation was applied for each chamber. Based on the measurements, there were totally 47 and 45 rainfall events during the growing season of 2006 and 2009, respectively. According to their durations, the half-hourly analyses were applied to 10 and 13 rainfall events in 2006 and 2009, respectively. We further defined the time for the soil respiration response to peak after a rainfall event as the peak time (Tpeak), and calculated the duration of the pulse (Tduration) before soil respiration decreased to 99% of the antecedent soil respiration. The rain pulse effects were estimated by the accumulative differences during this period. These analyses were conducted using Matlab (Mathworks, Natick, MA).

Linear regression analyses were used to evaluate the relationships between rain pulse response patterns (Tpeak, Tduration, pulse effect) of soil respiration, rain size and water status index (WSI) in the growing seasons of 2006 and 2009. The water status index (WSI) was defined as,(2)where STpre and SMpre are the soil temperature and moisture at 10 cm depth before a rainfall event. The index thus takes both rainfall event size and antecedent condition into consideration. The development of the WSI was based on the findings from previous studies, which have shown that larger rainfall event size [4] as well as drier [23] and warmer [22] pre-event soil conditions positively affect rain pulse effect. Growing-season total soil respiration was defined as the sum of daily soil respiration from May to October in each year.

Results

3.1. Growing-season total soil respiration and relative environmental factors

Soil respiration rates showed clear seasonal variations and reached maximum values in the middle of the growing seasons in both 2006 and 2009 (Fig. 1), with the growing-season total soil respiration as 187.2±16.47 g C m−2 in 2006 and 155.9±4.58 g C m−2 in 2009. Precipitation during the growing season (May to October) was slightly above the long-term average in 2006 (403 mm) but was far below the average in 2009 (168 mm) (Fig. 1). In both years, the heaviest rainfall occurred in the middle of the growing season (late July to early August; Fig 1). Accordingly, soil moisture fluctuated through the growing season in both years, with the highest soil moisture occurred in late July in both years (Fig. 1). Soil temperature had a pronounced seasonal dynamic in both years, with the highest temperature occurring in late July of 2006 and early August of 2009 (Fig. 1). Lower seasonal mean soil temperature (16.86°C) and higher soil moisture (11.26%) were found in 2006 than those of 2009 (17.14°C and 7.81%).

thumbnail
Figure 1. Seasonal patterns of precipitation (gray rectangle), soil moisture (black dashed line), soil temperature (black line) and soil respiration (gray line) during the growing seasons of 2006 (a) and 2009 (b).

https://doi.org/10.1371/journal.pone.0104217.g001

3.2. Pulse effects of soil respiration and its controlling factors

Equation (1) well fitted the temporal patterns of soil respiration following heavy rainfall events, with r2 ranging from 0.91 to 0.99 (P<0.05; Fig. 2a–f). In 2009, for example, daily soil respiration rates increased by 72–160% within 2 days following rainfall events (Fig. 2d–f). For small rainfall events, daily average of soil respiration could not capture the rapid responses of soil respiration to rainfall events. For example, similar to the predicted curves in Figure 3, the rainfall event effect on soil respiration peaked within several hours of the end of small rainfall events (Figure S1 in File S1). After reaching the peak, the soil respiration rate gradually returned to its antecedent level in 2.93 to 27.9 days after rainfall events of 0.4 mm to 31 mm, respectively (estimated from the fitted curve in Fig. 2). Both the magnitude and duration of the pulse effect on soil respiration (Tpeak and Tduration) were positively correlated with rainfall event size (Fig. 3a–c). A higher sensitivity of rainfall event effect to rainfall event size was found in 2009 than in 2006 (Fig. 3c). It could be ascribed to the higher sensitivities of Tpeak and Tduration to rainfall event size in 2009 than in 2006 (Fig. 3a, b). It suggested that rainfall event size alone was not a good indicator for the effect of rainfall event on soil respiration in this ecosystem. When compared against WSI instead of rainfall event size, however, there were no significant differences in regressive curves between 2006 and 2009 (P>0.1) (Fig. 3d–f). The values of Tpeak (about 3 days; r2 = 0.86, P<0.001) and Tduration (about 30 days; r2 = 0.94, P<0.001) of the pulse responses increased logarithmically with the WSI (Fig. 3d and e). Accumulative soil respiration during the pulse processes increased linearly with the WSI (Pulse effect  = 0.1169×WSI, r2 = 0.92, P<0.001, Fig. 3f). Based on this relationship, we calculated the pulse effect of each rainfall event on soil respiration, and found the pulse effect contributed 44.5% and 39.6% of the measured growing-season total soil respiration in 2006 and 2009, respectively.

thumbnail
Figure 2. The response of soil respiration to rain pulses under different rain sizes.

Open circles are the measured data and shown as mean ± standard deviation. Exponential equations in the form of y = y0+atebt are used to fit the data (represented by solid lines). Rain size is shown in each panel. The 0 in the x-axis represents the day before the rainfall event.

https://doi.org/10.1371/journal.pone.0104217.g002

thumbnail
Figure 3. The relationship between dependent variables including the peak time (Tpeak) and the duration (Tduration) of the pulses, and the pulse effects on soil respiration, and independent variables including rain size (a, b, c) and WSI (rain size × the ratio of antecedent soil temperature to soil moisture; d, e, f) in 2006 (solid circle) and 2009 (open circle), respectively.

https://doi.org/10.1371/journal.pone.0104217.g003

3.3. Long-term trend of the precipitation regime during 1953-2009

The mean precipitation during the growing season from 1953-2009 was 358 mm (Fig. 4b). The frequency of rainfall event occurrence decreased with increasing amount of rainfall (Fig. 4a). Long-term mean rainfall frequency with rainfall amount of 0–2 mm, 2–5 mm, 5–10 mm, 10–15 mm, 15–20 mm, 20–30 mm, 30–40 mm, 40–50 mm, and >50 mm per event was 48.4%, 21.8%, 14.0%, 7.3%, 3.8%, 2.8%, 1.0%, 0.4%, and 0.4%, respectively (Fig. 4a). We classified the duration of dry-spell into five groups (<5 days, 5–10 days, 10–20 days, 20–30 days, and >30 days). The long-term mean occurrence frequency of the dry-spells was 71.7%, 21.4%, 5.9%, 0.8%, and 0.3% for the five groups, respectively (Fig. 4c).

thumbnail
Figure 4. The frequency distribution of different class-categories of rainfall events (a) and dry-spell events (c) during the growing seasons of 2006 and 2009.

The long-term mean annual precipitations from 1953 to 2009 are also represented here by black columns. The insert panels show the precipitation during the growing season (b) and the mean dry-spell duration during the growing season (d) in 2006 and 2009.

https://doi.org/10.1371/journal.pone.0104217.g004

During the 57 years, no clear temporal trend of the total growing-season precipitation was found (Figure S2 in File S1). The 40–50 mm (r2 = 0.16, P = 0.061) and >50 mm (r2 = 0.23, P<0.05) rainfall events showed significant increasing tendencies over the years (Fig. S2k and l in File S1), while those of the smaller-sized rainfall events did not show any clear trend (Figure S2b–g in File S1). The mean dry-spell duration of the growing season increased from 1953 to 2009 (P<0.001; Figure S3a). During these 57 years, the frequency of dry spell events <5 days in duration significantly decreased (r2 = 0.28, P<0.001; Figure S3b in File S1), while the dry-spell events of 5–10 days (r2 = 0.13, P = 0.003) and 10–20 days (r2 = 0.11, P = 0.01) occurred more frequently (Figure S3c–d in File S1). The frequency of dry-spell events longer than 20 days did not change during this time period (Figure S3e–f in File S1).

Discussion

4.1. Factors controlling rain pulse on soil respiration

In this study, significant pulse responses of soil respiration were found after rainfall events (Fig. 2), which was consistent with the results reported in other arid ecosystems [18], [25]. The rain pulse induced soil respiration was 44.5% (83.3 g C m−2) and 39.6% (61.7 g C m−2) of the growing-season total soil respiration in 2006 and 2009, respectively. Since only few studies have quantified the contribution of rain pulse effect to growing-season total soil respiration, it is difficult to compare our results to those from other grasslands. However, some evidence show that the contribution of rain pulse to total soil respiration is lower in forest ecosystems. For example, Lee et al. [42] reported that the rainfall-induced soil CO2 release accounted for 16–21% of the annual soil respiration in a deciduous forest in Japan, whereas Yuste et al. [43] estimated that approximately 9–14% of the annual soil respiration in Belgian Campine region in Belgium was rainfall induced. The stronger rain pulse responses in grasslands could be the result of more severe water deficiency before rainfall events in grasslands and higher soil organic carbon content at the soil surface layer of grasslands than of shrublands or wood stands [36]. Therefore, the rain pulse effects on belowground C release is critical to ecosystem C balance in the semi-arid steppe ecosystem.

A rainfall event consists of at least three processes. Firstly, the percolating rainwater replaces the CO2 in soil pore spaces. Secondly, the rainwater activates microbial activity and induces microbe-respired CO2 in the shallow soil [44]. Thirdly, the rainwater also promotes the assimilation process by roots, which increases root respiration [3], [45]. These processes have been confirmed by the positive relationship found between the rainfall event effect on soil respiration and precipitation in previous studies through field manipulative experiments [4], [7], [18]. Similar trends have also been found in our study. During the two growing seasons of 2006 and 2009, we observed that a larger rainfall event induced a greater pulse effect on soil respiration (Fig. 2a–c). Rather than using the limited water and resources in the soil to maintain activity, most microbes may simply become dormant in the dry soil [46]. The sudden increase in soil water availability after rains can induce microbial cell lysis or the rapid mineralization of cytoplasmic solutes and release the mineralized product into the surrounding environment to dispose of its osmolytes, which have accumulated during the dry period [27], [47]. Consequently, rainfall event induced soil respiration largely comes from the decomposition of microbial cellular material. In addition, more rainfall means more water percolates to the rhizosphere and triggers root activity and respiration [3]. Therefore, the dependence of pulse effect and its duration of soil respiration on precipitation amount in this study can be explained by the greater increase in water availability under heavier rainfall event (Fig. 5).

thumbnail
Figure 5. The relationships between the rain size and the change in water availability after a rainfall event (ΔSM; the changes in the largest volumetric soil moisture after a rainfall event from the antecedent volumetric soil moisture) in 2006 (open circle) and 2009 (solid circle) with the coefficient (r2) and the significance level (P-value).

https://doi.org/10.1371/journal.pone.0104217.g005

Although the effects of rainfall events on the changes in water availability before and after rain were similar between 2006 and 2009 (Fig. 5), a significant inter-annual difference in response patterns of soil respiration to rainfall event was found (Fig. 2a–c). It indicates that some other factors rather than rainfall event size regulate the effect of rain pulse on soil respiration. For example, many studies [20], [23], [48] have found that the same amount of water addition had greater effects on soil respiration under drier antecedent soil condition. Soil temperature also has significant effect on soil respiration pulse [5], [24], [36,]. In addition, vegetation, which influences both water and C cycles, are important in regulating the effect of rain pulse on soil respiration [49], [50]. In this study, we introduced the WSI to predict the effect of rain pulse on soil respiration in two growing seasons with contrasting precipitation regimes (Fig. 3d–f). It should be noted that this empirical index does not explain the underlying mechanisms of the response of soil respiration to rainfall events, which have been studied by some hourly-scale stochastic models [37]. However, it highlights the importance of precipitation regime on the pulse response of soil respiration in semi-arid areas. Since the WSI was developed from only two growing-season measurements and based on five chambers, this index needs to be tested in other sites and years before it is used as an indicator for modeling and predicting rain pulse effect on soil respiration in semi-arid grasslands.

4.2. Precipitation regime shift and its influence on soil respiration

The long-term precipitation records showed no clear trend of total precipitation amount during a growing season but a higher frequency of heavy rainfall events in our study region (Figure S2 in File S1). It was consistent with the modeling predictions of greater precipitation intensity globally in the future [1], [51]. That is, more precipitation will fall in a given daily rainfall event, leading to more extreme rainfall events in the future. With similar total precipitation amount, higher frequency of heavy rainfall events means longer dry-spells during the growing season (Figure S3 in File S1). Such a shift in precipitation regime has been widely observed [40], [52], [53], although changes in precipitation amount would be greatly different among regions across the globe [2].

We found that a large proportion of the rain pulse effect on soil respiration was determined by the WSI, which considered the influences of rain size and antecedent soil temperature and moisture. The longer dry-spells are likely to lower the antecedent soil moisture. As a consequence, belowground C release is expected to be greater after a rainfall event compared to soils in rainfall regimes with a lower frequency of long dry-spells. The robustness of the WSI in explaining rain pulse effect on soil respiration needs to be examined in other types of grasslands. If the WSI is effective in most grassland ecosystems, it suggests not only total amount of annual precipitation but also the temporal distribution of rainfall events must be taken into consideration when predicting the belowground C release and evaluating C budget of an ecosystem undergoing climate change.

4.3. Implications for modeling soil carbon release

Currently, process-based ecosystem models are common tools for simulating and predicting the future states of terrestrial C cycle [2]. Most of these models are established at hourly or daily time scales without incorporating the immediate pulse responses of soil respiration to rainfall events. Therefore, rain pulse effect on soil respiration is one of the key sources of the uncertainty in the modeled ecosystem C cycle, especially in grasslands [8], [36]. In this study, we found that the rain pulse effect on soil respiration and its duration increased with the product of rain size and antecedent soil temperature-to-moisture ratio. However, the relationship between the WSI and rain pulse effect on soil respiration (y = 0.1169×WSI) as well as its duration (y = 30.13(1-e−0.021x)) was obtained from the data of only one grassland site during two growing seasons. The parameters in the regression model might vary with soil texture and soil organic matter content, which supplies C substrate for respiration [5], [20], [27]. Therefore, the applicability of the WSI we proposed in this study still needs to be tested in broader spatial scales and at more locations. Parameter information that constrains the relationship between the WSI and rain pulse effect on soil respiration could be useful for improving our ability to predict soil C cycle under future precipitation regimes.

There is evidence that CO2 from carbonates may [54] or may not [55] contribute significantly to total CO2 release when soils are moistened. Since the soil in our site is calcic, it is still unclear whether rain pulse effect will increase CO2 release from inorganic C sources. Future study may pay more attention to the pulse effect of rainfall events on not only the decomposition of soil organic matters but also CO2 release from inorganic C sources.

Conclusions

In this study, we presented a water status index (WSI) as a good indicator for predicting the effect of rain pulse on soil respiration in a temperate steppe in northern China. Long-term precipitation data showed a significant increase in frequency of large rainfall events and dry-spell duration with time. Although there are no measurements of soil temperature and moisture from the historical record to compare against measurements made in 2006 and 2009, the longer dry-spell may leads to larger antecedent ratio between temperature:moisture in the soils, and thus greater rain pulse effect on soil respiration. In this study, rain pulse effect contributed to about 39–44.5% of the growing-season total soil respiration, which is much larger than previous studies in this [4] and other [23] ecosystems. It indicates that the effect of rain pulses on soil respiration cannot be neglected in future climate-carbon cycle modeling. Future research could test the robustness of WSI in other regions before it is used as a simple approach to estimate the rain pulse effect on soil respiration in grassland ecosystems.

Supporting Information

File S1.

Supporting figures. Figure S1, Half-hourly soil respiration rate (SR, µmol m−2 s−1) under different small rainfall events. Rainfall size is shown in each panel. The 0 in the x-axis is the beginning time of a rainfall. Figure S2, The precipitation amount during the growing season (a) and the occurrence frequency of the nine class-categories of precipitation during the growing season (b–l) from 1953 to 2009. Figure S3, The mean dry-spell duration during the growing season (a) and the occurrence frequency of the five class-categories of the dry-spell duration during the growing season (b–f) from 1953 to 2009.

https://doi.org/10.1371/journal.pone.0104217.s001

(DOCX)

Acknowledgments

We thank Dr. Bazartseren Boldgiv and two anonymous reviewers for their valuable comments and suggestions. Thanks, Ms. Jin Chen, for polishing the English language of an early version of this manuscript.

Author Contributions

Conceived and designed the experiments: SC LY. Analyzed the data: LY SC. Contributed to the writing of the manuscript: LY SC. Conducted the filed measurement: LY. Provided constructive suggestions on data analysis: JX YL. Provided substantial comments and revised the early version of manuscript: JX YL.

References

  1. 1. Dore MHI (2005) Climate change and changes in global precipitation patterns: What do we know? Environment International 31: 1167–1181.
  2. 2. IPCC (2007) Climate change 2007: the physical science basis. In: Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, Tignor M, Miller HL, editors, Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press. pp. 30–33.
  3. 3. Huxman TE, Snyder KA, Tissue D, Leffler AJ, Ogle K, et al. (2004) Precipitation pulses and carbon fluxes in semiarid and arid ecosystems. Oecologia 141: 254–268.
  4. 4. Chen SP, Lin GH, Huang JH, Jenerette GD (2009) Dependence of carbon sequestration on the differential responses of ecosystem photosynthesis and respiration to rain pulses in a semiarid steppe. Global Change Biology 15: 2450–2461.
  5. 5. Kim DG, Vargas R, Bond-Lamberty B, Turetsky MR (2012) Effects of soil rewetting and thawing on soil gas fluxes: a review of current literature and suggestions for future research. Biogeosciences 9: 2459–2483.
  6. 6. Liu R, Cieraad E, Li Y (2013) Summer rain pulses may stimulate a CO2 release rather than absorption in desert halophyte communities. Plant and Soil 373: 799–811.
  7. 7. Liu XZ, Wan SQ, Su B, Hui DF, Luo YQ (2002) Response of soil CO2 efflux to water manipulation in a tallgrass prairie ecosystem. Plant and Soil 240: 213–223.
  8. 8. Lee X, Wu HJ, Sigler J, Oishi C, Siccama T (2004) Rapid and transient response of soil respiration to rain. Global Change Biology 10: 1017–1026.
  9. 9. Ma S, Baldocchi DD, Hatala JA, Detto M, Yuste JC (2012) Are rain-induced ecosystem respiration pulses enhanced by legacies of antecedent photodegradation in semi-arid environments? Agricultural and Forest Meteorology 154–155: 203–213.
  10. 10. Högberg P, Nordgren A, Buchmann N, Taylor AFS, Ekblad A, et al. (2001) Large-scale forest girdling shows that current photosynthesis drives soil respiration. Nature 411: 789–792.
  11. 11. Davidson EA, Janssens IA (2006) Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature 440: , 165–173.
  12. 12. Wan S, Norby RJ, Ledford J, Weltzin JF (2007) Responses of soil respiration to elevated CO2, air warming, and changing soil water availability in a model old-field grassland. Global Change Biology 13: 2411–2424.
  13. 13. Griffiths E, Birch HF (1961) Microbiological Changes in Freshly Moistened Soil. Nature 189: 424.
  14. 14. Kieft TL, Soroker E, Firestone MK (1987) Microbial biomass response to a rapid increase in water potential when dry soil is wetted. Soil Biology and Biochemistry 19: 119–126.
  15. 15. Borken W, Xu YJ, Davidson EA, Beese A (2002) Site and temporal variation of soil respiration in European beech, Norway spruce, and Scots pine forests. Global Change Biology 8: 1205–1216.
  16. 16. Schimel JP, Gulledge JM, Clein-Curley JS, Lindstrom JE, Braddock JF (1999) Moisture effects on microbial activity and community structure in decomposing birch litter in the Alaskan taiga. Soil Biology and Biochemistry 31: 831–838.
  17. 17. Chen SP, Lin GH, Huang JH, He M (2008) Responses of soil respiration to simulated precipitation pulses in semiarid steppe under different grazing regimes. Journal of Plant Ecology-UK 1: 237–246.
  18. 18. Sponseller RA (2007) Precipitation pulses and soil CO2 flux in a Sonoran Desert ecosystem. Global Change Biology 13: 426–436.
  19. 19. Vargas R, Collins SL, Thomey ML, Johnson JE, Brown RF, et al. (2012) Precipitation variability and fire influence the temporal dynamics of soil CO2 efflux in an arid grassland. Global Change Biology 18: 1401–1411.
  20. 20. Cable JM, Ogle K, Williams DG, Weltzin JF, Huxman TE (2008) Soil texture drives responses of soil respiration to precipitation pulses in the Sonoran Desert: Implications for climate change. Ecosystems 11: 961–979.
  21. 21. Bowling DR, Grote EE, Belnap J (2011) Rain pulse response of soil CO2 exchange by biological soil crusts and grasslands of the semiarid Colorado Plateau, United States. Journal of Geophysical Research -Biogeosciences 116: G03028
  22. 22. Carbone MS, Still CJ, Ambrose AR, Dawson TE, Williams AP, et al. (2011) Seasonal and episodic moisture controls on plant and microbial contributions to soil respiration. Oecologia 167: 265–278.
  23. 23. Wu HJ, Lee X (2011) Short-term effects of rain on soil respiration in two New England forests. Plant and Soil 338: 329–342.
  24. 24. Jager G, Bruins EH (1975) Effect of Repeated Drying at Different Temperatures on Soil Organic-Matter Decomposition and Characteristics, and on Soil Microflora. Soil Biology and Biochemistry 7: 153–159.
  25. 25. Smart DR, Penuelas J (2005) Short-term CO2 emissions from planted soil subject to elevated CO2 and simulated precipitation. Applied Soil Ecology 28: 247–257.
  26. 26. Vargas R, Allen MF (2008) Environmental controls and the influence of vegetation type, fine roots and rhizomorphs on diel and seasonal variation in soil respiration. New Phytologist 179: 460–471.
  27. 27. Fierer N, Schimel JP (2003) A proposed mechanism for the pulse in carbon dioxide production commonly observed following the rapid rewetting of a dry soil. Soil Science Society of America Journal 67: 798–805.
  28. 28. Riveros-Iregui DA, Emanuel RE, Muth DJ, McGlynn BL, Epstein HE, et al. (2007) Diurnal hysteresis between soil CO2 and soil temperature is controlled by soil water content. Geophysical Research Letters 34: L17404
  29. 29. Liu WX, Zhang Z, Wan SQ (2009) Predominant role of water in regulating soil and microbial respiration and their responses to climate change in a semiarid grassland. Global Change Biology 15: 184–195.
  30. 30. Yan LM, Chen SP, Huang JH, Lin GH (2011) Water regulated effects of photosynthetic substrate supply on soil respiration in a semiarid steppe. Global Change Biology 17: 1990–2001.
  31. 31. Liu BH, Xu M, Henderson M, Qi Y (2005) Observed trends of precipitation amount, frequency, and intensity in China, 1960–2000. Journal of Geophysical Research-Atmospheres 110: D08103
  32. 32. Fierer N, Schimel JP (2002) Effects of drying-rewetting frequency on soil carbon and nitrogen transformations. Soil Biology and Biochemistry 34: 777–787.
  33. 33. Miller AE, Schimel JP, Meixner T, Sickman JO, Melack JM (2005) Episodic rewetting enhances carbon and nitrogen release from chaparral soils. Soil Biology and Biochemistry 37: 2195–2204.
  34. 34. Knapp AK, Fay PA, Blair JM, Collins SL, Smith MD, et al. (2002) Rainfall variability, carbon cycling, and plant species diversity in a mesic grassland. Science 298: 2202–2205.
  35. 35. Harper CW, Blair JM, Fay PA, Knapp AK, Carlisle JD (2005) Increased rainfall variability and reduced rainfall amount decreases soil CO2 flux in a grassland ecosystem. Global Change Biology 11: 322–334.
  36. 36. Xu LK, Baldocchi DD, Tang JW (2004) How soil moisture, rain pulses, and growth alter the response of ecosystem respiration to temperature. Global Biogeochemical Cycles 18: GB4002.
  37. 37. Daly E, Oishi AC, Porporato A, Katul GG (2008) A stochastic model for daily subsurface CO2 concentration and related soil respiration. Advances in Water Resources 31: 987–994.
  38. 38. Ni J (2002) Carbon storage in grasslands of China. Journal of Arid Environments 50: 205–218.
  39. 39. Lee R, Yu FF, Price KP, Ellis J, Shi PJ (2002) Evaluating vegetation phenological patterns in Inner Mongolia using NDVI time-series analysis. International Journal of Remote Sensing 23: 2505–2512.
  40. 40. Gong DY, Shi PJ, Wang JA (2004) Daily precipitation changes in the semi-arid region over northern China. Journal of Arid Environments 59: 771–784.
  41. 41. Paulo AA, Pereira LS (2006) Drought concepts and characterization: Comparing drought indices applied at local and regional scales. Water International 31: 37–49.
  42. 42. Lee MS, Nakane K, Nakatsubo T, Mo WH, Koizumi H (2002) Effects of rainfall events on soil CO2 flux in a cool temperate deciduous broad-leaved forest. Ecology Research 17: 401–409.
  43. 43. Yuste JC, Janssens IA, Ceulemans R (2005) Calibration and validation of an empirical approach to model soil CO2 efflux in a deciduous forest. Biogeochemistry 73: 209–230.
  44. 44. Sharkhuu A, Plante AF, Enkhmandal O, Casper BB, Helliker BR, et al. (2013) Effects of open-top passive warming chambers on soil respiration in the semi-arid steppe to taiga forest transition zone in Northern Mongolia. Biogeochemistry 115: 333–348.
  45. 45. Reynolds JF, Kemp PR, Ogle K, Fernandez RJ (2004) Modifying the ‘pulse–reserve’ paradigm for deserts of North America: precipitation pulses, soil water, and plant responses. Oecologia 141: 194–210.
  46. 46. Boot CM, Schaeffer SM, Schimel JP (2013) Static osmolyte concentrations in microbial biomass during seasonal drought in a California grassland. Soil Biology and Biochemistry 57: 356–361.
  47. 47. Schimel J, Balser TC, Wallenstein M (2007) Microbial stress-response physiology and its implications for ecosystem function. Ecology 88: 1386–1394.
  48. 48. Yan LM, Chen SP, Huang JH, Lin GH (2010) Differential responses of auto- and heterotrophic soil respiration to water and nitrogen addition in a semiarid temperate steppe. Global Change Biology 16: 2345–2357.
  49. 49. Kim DG, Mu S, Kang S, Lee D (2010) Factors controlling soil CO2 effluxes and the effects of rewetting on effluxes in adjacent deciduous, coniferous, and mixed forests in Korea. Soil Biology and Biochemistry 42: 576–585.
  50. 50. Vargas R, Detto M, Baldocchi DD, Allen MF (2011) Multiscale analysis of temporal variability of soil CO2 production as influenced by weather and vegetation. Global Change Biology 16: 1589–1605.
  51. 51. Meehl GA, Arblaster JM, Tebaldi C (2005) Understanding future patterns of increased precipitation intensity in climate model simulations. Geophysical Research Letters 32: L18719.
  52. 52. Schmidli J, Frei C (2005) Trends of heavy precipitation and wet and dry spells in Switzerland during the 20th century. International Journal of Climatology 25: 753–771.
  53. 53. New M, Hewitson B, Stephenson DB, Tsiga A, Kruger A, et al. (2006) Evidence of trends in daily climate extremes over southern and west Africa. Journal of Geophysical Research-Atmospheres 111: D14102
  54. 54. Tang JW, Baldocchi DD (2005) Spatial-temporal variation in soil respiration in an oak-grass savanna ecosystem in California and its partitioning into autotrophic and heterotrophic components. Biogeochemistry 73: 183–207.
  55. 55. Inglima I, Alberti G, Bertolini T, Vaccari FP, Gioli B, et al. (2009) Precipitation pulses enhance respiration of Mediterranean ecosystems: the balance between organic and inorganic components of increased soil CO2 efflux. Global Change Biology 15: 1289–1301.