Loading

Establishment and Production of Switchgrass Grown for Combustion: A Review

Review Article | Open Access

  • 1. Stockbridge School of Agriculture, University of Massachusetts, USA
+ Show More - Show Less
Corresponding Authors
Stockbridge School of Agriculture, University of Massachusetts, USA
Abstract

Switchgrass (Panicum virgatum L.) is a C4 -grass with deep fibrous root systems indigenous to North America. In recent years switchgrass has been considered to be a “model” energy crop due to its high productivity, perenniality, and adaptability to various sites and soils. This paper specifically reviews published works on the effect of cultural management practices on switchgrass establishment, biomass production and composition, dynamic of nutrient and non-structural carbohydrates (NSC) translocation from above-ground to roots and nitrogen-use efficiency (NUE).

Citation

Hashemi M, Sadeghpour A (2013) Establishment and Production of Switchgrass Grown for Combustion: A Review. Int J Plant Biol Res 1(1): 1002.

Keywords

•    Biomass
•    Establishment
•    Harvesting
•    Nitrogen-use efficiency
•    Switchgrass

INTRODUCTION

In recent years, notable interest has been paid to biomassbased energy production due to economic and environmental issues related to fossil fuel [1].Use of grain corn (Zea mayes L.) as the common feedstock for ethanol production has raised serious concerns about its sustainability. These concerns are mainly related to environmental pollution due to increased soil erosion and high agricultural inputs including chemical fertilizers and herbicides. Therefore use of perennial plant species (grasses and woods) which are more environmentally friendly sources of bioenergy production has gained more and more attention [2]. A ten-year study that began in the 1980’s at Oakland Ridge National Laboratory identified switchgrass as an ideal species for bioenergy productiondue to variety of its desirable characteristics [3]. Consequently, dedicated research effort over the last thirty years, has led to significant progress in developing switchgrass as a biofuel crop [4]. The ultimate use of switchgrass is commonly either ethanol or heat [5]; however, when cultivation land is limited, energy production through combustion seems more feasible [6]. In this article the challenges associated with establishment, survival, and production of switchgrass grown for combustion are discussed.

SWITCHGRASS PLANT OVERVIEW

Switchgrass is a warm-season (C4), sod-forming perennial tall grass native to North America [7]with deep fibrous roots which can reach up to 3 m deep [8]. The species has been evolving since approximately two million years ago and its dispersal from tropical regions to Central and North America created an extensive genotypic variation among the crop species leading to high adaptation of switchgrass to a wide range of growing conditions [9]. Latitudinal differences are most responsible for variation among switchgrass populations. Latitude of origin has been reported to have a significant impact on productivity, survival, and adaptation traits of switchgrass [10,11]. In 1966, Porter [12] categorized switchgrass populations between two distinct ecotypes; “upland” and “lowland.” Lowland ecotypes occur in lower hydric conditions in lower latitudes, whereas upland varieties occur in drier, elevated conditions and are more common at higher latitudes [13]. Lowland ecotypes are more tolerant of wet conditions than upland types and grow taller and faster, but are more sensitive to drier conditions [14]. The leaves of lowland switchgrass are bluish-green and coarser and thicker than upland varieties. Additionally, the ligules are longer and the panicles are larger than upland types [15]. Upland ecotypes have thin stems, are generally less productive than the lowland varieties, often grow in a bunch form and are adapted to dry conditions [16]. Although lowland ecotype is less tolerant to dry conditions, the extensive root systems of switchgrass allow for both ecotypes to be more drought tolerant than other herbaceous crops such as Miscanthus (Miscantus giganteum L.). Elberson et al. [17] determined that latitudinal differences were the main factor influencing adaptability, when southern varieties had higher yields in the north than northern varieties. When grown too far north however, southern varieties could be winter-killed [3]. In general, northern ecotypes have a longer winter dormant period with better winter survival than southern ecotypes when grown at the same latitude [18]. Conversely, planting northern varieties in southern locations does not necessarily maximize the yield because these varieties cease growth sooner in the fall due to their adaption to shorter growing season [19]. Figure 1 illustrates biomass yield differences between upland and lowland cultivars [20]. Among lowland ecotypes, the most productive cultivars were Alamo, SL941, SL931, Kanlow, NL942 and SL932 with average biomass production of 12.2 to 14.8 Mg ha-1. Within upland ecotypes, Cave-in-Rock, NE Late, HDMDC3, Late-Synthetic-HY, Shelter, and NU94 were the highest yielding cultivars with median yield of annual biomass production ranged from 9.6 to 11.4 Mg ha-1.

High adaptation to various sites and soils [21] plus high productivity with low chemical input (i.e. nitrogen fertilizer, herbicides and pesticides, etc.) [22], perenniality [23] as well as feasibility of harvest with conventional hay-making equipment [24] have been general criteria for the selection of switchgrass as a promising dedicated energy crop [21].The ability of switchgrass to positively influence the environment by sequestering carbon (C), reducing soil and wind erosions, and increasing wildlife habitat has also been considered and well documented [5,25,3].

ESTABLISHMENT MANAGEMENT

One of the important challenges in switchgrass production is seedling establishment [23,4]. Similar to many warm-season perennial grasses, switchgrass has been known to be difficult or slow to establish [23,26,27]. Poor establishment in the planting year directly relates to reduced stand vigor and yield in succeeding years and limits large scale crop adoption [10,28,29]. It is estimated that a stand failure costs growers over $300 ha-1 [30].

Switchgrass initially allocates energy to establishing an extensive root system in the first and second year and will consequently only reach 33 and 66% of its maximum production capacity, respectively [31]. Due to the allocation of energy to the development of root structures, switchgrass will not reach its full yield potential until the third year [32]. This extended establishment time has dissuaded many growers and entrepreneurs from planting switchgrass given the lack of financial return in the first two years; however with proper planning, switchgrass can be profitable endeavor for growers.

Establishment of switchgrass specifically in the establishing year can be influenced by several factors including high seed dormancy and weed pressure, improper planting technique or seedbed preparation, and adverse environmental conditions [3,23,33].

Seed dormancy

Seed dormancy is one of the major challenges in establishment of switchgrass [28]. Switchgrass seed has been proven to be highly dormant at seed dispersal [34,35,36]. Innate seed dormancy can be caused by many chemical or physical inhibition mechanisms; however, it is most often due to the immaturity of the seed embryo at dispersal [37,38]. Chemical inhibition is caused by hormones that restrict germination [38] whereas; physical inhibition is caused by seed coat barrier [39]. One strategy to increase germination rates for maximum stand establishment is to reduce seed dormancy [40]. Dormancy reduction can be achieved through various methods. Two common approaches are stratification and after-ripening [41]. Studies concluded that stratification or a wet pre-chilling treatment at 5°C for two or more weeks reduced dormancy rates [42,43]. Averaged over two Cave-In-Rock seedlots, Shen et al. [41] found that stratification at 5°C for 14 days increased germination from 7% to 75%. Zhang and Maun [38] also found that germination rates could be increased from 3% to anywhere from 88-98% by scarification of the seed coat. Although this method was successful, in a review article, Parrish and Fike [3] stated that seed priming, scarification and hormonal treatments may not be applicable strategies on large-scale switchgrass production. One seed dormancy-breaking technique that is more feasible for large-scale production is after-ripening, storage of seeds for one or more years in a warm environment, which has shown positive practical effects on the reduction of dormancy in switchgrass [44].

Sowing rate

Variable germination rates of switchgrass due to seed dormancy can confound determination of sowing rate [14]. Several studies have developed, various planting rate recommendations have been made based on different calculation methods. Whether based on mass per area or number of “pure live seeds” per area, there have been many points of confusion regarding this matter [3]. Pure live seed (PLS) refers to seed that is viable, including both dormant and non-dormant seeds[10]. In a standard germination test [45], results would be lower than in a viability test for PLS because dormant seeds will not necessarily germinate [46]. Seed distributors often test their seeds for viability (PLS), germination rate, weed seed contaminations and inert matter and include the test results on the seed packaging. Using the distributor’s test results for PLS (%) or germination (%) to calculate planting rates will lead to an inaccurate planting rate [47]. Due to reduction in dormancy rates over time, current germination percentages do not necessarily correspond with supplied information. Conversely, seed testing laboratories will present inflated test data collected from controlled environment that do not accurately represent the stressed conditions that might occur in the field. In summary, the use of seed distributors’ test results for determination of sowing rate should be avoided. Forberg et al. [48] concluded that it is more practical to implement a vigor test and then compensate for restricted germination by adjusting sowing rates.

Precise planting rates are crucial for a successful and economical planting of switchgrass as a bioenergy crop [29]. A low stand frequency will limit yield and too high of a stand frequency will waste seed [49]. The average recommended planting rate is 4 to 10 kg ha-1 PLS [33,36,50]. Alternatively, recommendations have been made based on number of established plants per m-2. Teel et al. [36] recommended 20 plants per m-2 as an adequate established stand for bioenergy usage; however, it is difficult to plant at a rate targeted for number of established plants per area. Forberg [14] found 30-50% seedling mortality after emergence across four varieties (Blackwell, Carthage, Cave-in-Rock, and Dacotah) grown in Massachusetts. He also observed higher seedling mortality with higher seeding rates. Stand densities of 278 plants m-2 from 600 PLS m-2 could be achieved [3]. Ultimately, desired stand frequency or density is the principle consideration for the determination of planting rates. Vogel and Masters [49] designed a frequency grid with which stand density of switchgrass could be determined. In their previous switchgrass establishment research, frequency-grid-measured switchgrass stands of 40 to 50% or greater indicated a successful stand, frequencies between 25 to 50% were marginal to adequate, and frequencies <25% indicated partial stands that need replanting [47,51]. Mitchell and Schmer [47] reported that in most cases, poor seed quality resulted in poor stand establishment that required re-planting.

Other factors that affect the establishment of switchgrass include soil preparation and seeding methods, seed placement, planting date, weed control, and environmental conditions [23,52].

Seeding methods

Methods of seedbed preparation for planting switchgrass typically include: conventional and no-till planting into killed sods or bare soil [3]. Although several reports have indicated the preference of conventionally tilled seedbeds over no-till planting [36,53,54], no-till planting of switchgrass has also been proven to be useful in some circumstances [55]. There is limited information regarding the suitability of various seedbed preparations for switchgrass cultivation in different conditions [3]. McKenna et al. [56] and Teel et al. [36] suggested that planting into an herbicide-killed sod is possible with proper equipment, but they also stated that switchgrass stands planted using this method may be reduced compared with switchgrass stands planted into conventionally tilled seedbeds. Similarly, several researchers [26,53,54] suggested that switchgrass planted through direct drilling into killed sod was a less reliable method when compared with conventional tillage. In another approach, Monti et al. [23] showed that establishment of switchgrass was enhanced when conventionally prepared seedbeds were rolled or compacted after seeds were broadcasted. It is now well documented that switchgrass emergence increases greatly in a firm seedbed bed [23,26,57,58]. Venturi et al. [58] showed greatest germination in two varieties of switchgrass in well-tilled soil that was compacted before and after planting. They found lowest germination in tilled treatments without any compaction. Sadeghpour et al. [57], similarly reported that greatest germination rate, stand density, and biomass production was found when switchgrass was compacted two times after planting either with a roller or a cultipacker. In dry conditions, increasing seed-soil contact could also enhance germination through higher available moisture to the seeds. In contrast, other reports indicated no yield advantage from conventional tillage over no-till planting. For example, Rehm [59] found no switchgrass yield difference between no-till and conventional planting methods. King et al. [60] compared no-till with conventional planting of switchgrass at two locations in Nebraska and found that the yield advantage of one tillage system over the other was dependant on season and location. In a series of studies in Tennessee a 50 to 150% increase in switchgrass seedlings in a no-till system compared with a conventional seedbed preparation was found [61]. Sadeghpour et al. [27] found significant advantage of no-till planting over conventional tillage when precipitation was low during the growing season. In the same study, they used cereal cover crops, which are known to be fast growing and able to suppress weeds and provide N for the subsequent crop[62,63] to control weeds and enhance switchgrass establishment and found oat as the most effective cover crop for switchgrass establishment [27]. It could be concluded that the advantage of no-till planting of switchgrass over conventional tillage was partly due to soil and water conservation and also to the potential for earlier planting [3,55]. It is yet to be determined which planting method should be preferred due to various results in different locations.

Depth of Planting

Depth of seed placement is critical in emergence and the establishment of switchgrass [3]. In general, planting depths of 1 to 2 cm have been recommended to growers based on several studies [10,26,33,36]. Newman and Moser [64] found no significant difference between switchgrass emergence in plantings depths at 1.5 and 3 cm. However, they observed a 40% emergence reduction when they increased the sowing depth to 4.5 cm. It has also been suggested the emergence can affected by soil texture in conjunction with planting depth and moisture level. Aiken and Springer [65] found that soil texture and seed size among switchgrass cultivars had a greater effect on emergence than differences in planting depths within < 2 cm. Planting depths < 1 cm in sandy soils may result in low seedling survival under drought stress condition. Conversely, seedlings established in a clay soil at the same depths showed high survival at the same level of water stress [66]. In a recent greenhouse study, Berti and Johnson [10] observed significant differences on switchgrass emergence between surface planting (0 cm) and planting at the depth of 1.3 cm; however, did not find any significant differences in planting depths of 1.3 to 6.4 cm. In a field study the same authors found silty-clay soil as a more suitable media for switchgrass emergence compared with fine-silty and coarse-loamy soils in North Dakota, USA. In a greenhouse study, we also found a shallow planting < 3 cm could be suitable for switchgrass planting.

Seed size is also a factor in seedling emergence and vigor [3]. In several studies, larger seeds produced more vigorous seedlings in a shorter duration than smaller seeds [65,66]. In contrasts, Zhang and Maun [38] found no difference after eight weeks between seedlings from small or large seeds.

Date of Planting

Successful establishment of switchgrass acquires a sufficient stand that will maximize yield in subsequent years [67]. Planting dates can vary from November to July depending on several factors including geographical region; weed control methods; soil temperature; and rainfall patterns [3,23,58,68]. In warmer climates with longer growing seasons, switchgrass can be planted earlier than in cooler climates. However, planting early in the spring in most climates will cause slower seedling emergence than later plantings due to extreme temperature fluctuation and weed competition [14]. Optimal soil temperature for germination of a wide range of switchgrass cultivars have been suggested to be between 27-30°C However, according to Hsu et al. [68], a soil temperature of 20°C is sufficient for switchgrass emergence and growth. In a field study in Missouri, researchers found emergence to be more rapid at later planting dates in a set of treatments from April to June [68,69]. Similarly, in Massachusetts, we found faster emergence in June and July plantings compared with November and May. However, earlier-planted switchgrass was taller, and had more advanced root systems. In agreement with our findings, in Nebraska, Smart and Moser [42] found much larger seedlings and more vigorous stands in the earlier planting treatments spanning from March to late May. When comparing fall and spring plantings in a Mediterranean climate, Monti et al. [23] found slightly more emergence in spring plantings. Planting in a cool season could benefit seedling establishment by breaking dormancy in seeds by stratification. Hsu et al. [70] found that germination of dormant seeds increases in cool planting conditions. In several other studies, spring plantings of highly dormant seed yielded greater germination than later plantings; [36,71] however, this directly depends on the weather conditions. We found that in a mild winter with low amount of precipitation, emergence did not increase whereas; a cold and wet winter resulted in significant increase in switchgrass germination [Sadeghpour, unpublished data]. When rainfall proliferates in the spring, early plantings of switchgrass could be successful with proper weed control. But in many climates, weed pressure is high in early spring given warm temperature and increased rainfall [26,33]. Weed pressure in the establishment year can be reduced by avoiding planting at a time when weed emergence is high. Many annual weed species have a short period of emergence in the spring; therefore, delaying planting by two weeks could have positive effects on establishment [72]. In northern climates weed pressure is highest in the spring and thus planting should be delayed until early summer. There must be a balance between a delayed planting date for weed pressure avoidance while still allowing for enough growing season for adequate stand establishment [72].

Weed Control

A relatively small seed size, high dormancy rate, and slow germination often makes switchgrass a weak competitor with many summer annual grass and broadleaf weeds [73,74]. As a result, crop establishment and early growth is often delayed [29]. A poor switchgrass stand during the seeding year can limit yield and large scale crop adoption [10,28,29]. Weeds reduce yields of switchgrass by competing for nutrients, water, light, and space [75,76,77]. Additionally, some weed species produce toxins and growth inhibitors that can cause negative effects on switchgrass [78]. Switchgrass seedlings grow slowly in the first several months and can be out-competed by fast growing annual weeds [27]. Additionally, a major obstacle in weed management inperennial grasses is the lack of registered herbicides approved for this use [3]. In order to avoid stand failure, weed management must be a primary consideration in the establishment year of switchgrass [67]. Cool-season grassy weeds that germinate in cooler temperatures are most threatening to newly emerging switchgrass seedlings. Hsu and Nelson [68] found that crabgrass (Digitaria sanguinalis L.), a very problematic weed species, can grow more rapidly than switchgrass at equal temperature. Crabgrass produced up to 20 times more biomass per seedling than switchgrass when grown side by side. In our field trials in Massachusetts, crabgrass was also the most problematic weed in establishment of switchgrass which resulted in a significant reduction in stand density and yield [27].

The most effective weed management strategy in the establishment year could be herbicide application [29]. Efficacy of weed pressure reduction through herbicide application has been documented by several researchers [27,29,73,79]. For conventionally-tilled plantings, many studies have shown success with pre-emergent triazine herbicides, notably atrazine [6-chloro-N-ethyl-N’-(1-methylethyl)- 1,3,5-triazine-2.4-dimine] [79,29,27]. Switchgrass is one of the most tolerant grass species to atrazine [72]. Atrazine effectively controls many annual weed species when grown with perennial warm-season grasses [80,81,56]. Problematic weeds such as crabgrass, fall panicum (Panicum dichotmiflorum L.), foxtail species (Setaria spp.), and barnyardgrass (Echinochloa crus-galli L.) are less susceptible to atrazine treatments and require additional herbicide treatments for effective control. With similar growth habits to switchgrass, the control of these grassy weeds is crucial to avoid detriment to switchgrass stands [82]. Sadeghpour et al. [27] found sufficient weed control by using a combination of 1.1kg a.i. ha-1 atrazine and 0.37 kg a.i. ha-1quinclorac (3, 7-dichloro-8-quinolinecarboxylic acid). Quinclorac (Paramount) is highly effective at controlling annual warm-season grassy weeds as well as some broad leaf weeds and has recently been registered for use in switchgrass production [73,74]. Mitchell et al. [29] reported that a combination of quinclorac and atrazine provided satisfactory weed control for establishing both lowland and upland switchgrass cultivars in the Central and Northern Great Plains. Boydsten et al. [73] reported switchgrass yield and stand loss as a result of post-emergent application of quinclorac however, application of this herbicide in controlling grasses has been found to be very effective [27,73,79]. In a study at Wisconsin, Miesel et al. [83] reported that a mixed application of imazapic (±)-2- [4,5-dihydro-4-methyl-4-(1-methylethyl)- 5-oxo-1H-imidazol-2-yl]-5-methyl-3-pyridinecarboxylic acid] and glyphosate [N-(phosphonomethyl) glycine] at 0.07 kg a.i. ha-1 provided the best grassy weed suppression and resulted in the highest yield compared with different rates of glyphosate alone (1.12 kg a.i. ha-1) or in combination with 2,4- D [(2,4-dichlorophenoxy)acetic acid] at 1.06 kg a.i. ha-1. Kering et al. [84] studied the effect of various herbicides on switchgrass establishment and reported that when quinclorac was mixed with foramsulfuron [1-(4,6-dimethoxypyrimidin-2-yl)-3-(2- dimethylcarbamoyl-5-formamidophenyl-sulfonyl)urea)] and pendimethalin(3,4-Dimethyl-2,6-dinitro-N-pentan-3-yl-aniline) efficacy of weed control was more than 70% and switchgrass establishment was improved 13 to 26% compared to untreated control, however, their findings suggest that establishment was marginal and should be improved.

Broadleaf weeds in switchgrass can be controlled by an application of dicamba (3,6-dichloro-o-anisic acid) and 2,4-D [85]. In a recent study, Curran et al. [79] reported that a broadspectrum post-emergence application of atrazine, quinclorac, dicamba and 2,4-D significantly reduced the weed pressure in the establishment year of switchgrass. Findings of Sadeghpour et al. [27] are in line with earlier reports by Curran et al. [74,79] showing the effectiveness of a broad-spectrum application of atrazine, quinclorac, dicamba and 2,4-D. Further research is needed on herbicide application rates and their effect on switchgrass varieties.

One of the modern approaches to increase the success of herbicide application, reduce herbicide injury and enhance switchgrass establishment is seed safening [86]. Herbicide safeners can prevent herbicide damage of specific crops by reducing the binding abilities of molecules to affect target sites of plants [86]. This can be accomplished through safener-induced stimulation of herbicide catabolizing enzymes, or by safener-enhanced metabolism of herbicides to immobile metabolites [87,86]. Previously, seed safeners were proven to be effective in protecting several forage plants including sorghum (Sorghum biocolor L. Moench), perennial ryegrass (Lolium perenne L.), and sand bluestem (Andropogonhallii hack) from herbicide injury. To reduce the injury of switchgrass from pre-emergence application of metolacholor[2- chloro-N-(2-ethyl-6-methylphenyl)-N-(methoxy-1-1methyllethyl) acetamide], Rushing et al. [86] used two methods of seedsafening with fluxofenim (coating vs. controlled hydration). They reported that the controlled hydration (comination of 25, 50, or 100% fluxofenim) resulted in greater yields compared with the coating technique. Before this attempt, Butler et al. [88] was failed to safen switchgrass seeds in greenhouse experiments using fluxofenim.

In no-till plantings of switchgrass, weeds can be controlled effectively with a non-selective herbicide most notably glyphosate before the emergence of switchgrass [67].

As discussed earlier, planting date has a significant effect on weed pressure. Delaying seeding to allow weed emergence before final seed bed preparation will reduce weed pressure [77]. Curran et al. [79] found that delaying the planting until late June, resulted in weed pressure reduction.

PRODUCTION MANAGEMENT

Harvest

Harvesting strategy is dependent upon expected yield, quality and stand maintenance [89]. Frequency and time of harvest are the most important harvest management practices followed by cutting height [90].

Switchgrass harvesting frequency ranges from single-cut to multiple cuttings. Multiple harvests have been a viable strategy for forage agronomists to increase annual yield [3]. Commonly, after plants reachtheir maximum biomass, they can be harvested before the end of a growing season to allow for re-growth and increase total yield; however, many studies on switchgrass have shown multiple harvests results in yield reduction in succeeding years [3,11,91,92,93]. Madakadze et al. [91] found that a single end-of-season harvest was a more sustainable management practice compared with two or three cuttings. In the south-central USA, Sanderson et al. [11] reported that a single harvest at approximately 260 days of year provided the maximum biomass yield. They also concluded that multiple harvests (three or more) reduced yields over a 4-yr study. Generally, mid-summer harvests remove N and other nutrients from the shoots which would otherwise be translocated into the roots and crowns for successful re-growth in the following year. In a 5-yr study in Tennessee, Reynolds et al. [92] found no yield advantage of two-harvesting system (mid-summer and late-October) over a singlecut in late-October. Similarly, in a trial comparing numbers of harvests, Smart et al. [93] reported the benefits of a single harvest with respect to yield production. They found higher yields in onecut compared with total biomass produced by a three cutting system. An additional reason for yield reduction in long term studies is tiller density reduction [93,94]. Researchers in Virginia concluded that only a single or at most two-cut management could be appropriate to maximize biomass output [3,95,96].

In addition to harvest frequency time of harvest also influences switchgrass production [22,97] and perhaps is the most important harvest management practice [47]. Recommendations for the ideal time to harvest switchgrass to produce consistent maximum yield varies from site-to-site. A Mid-September harvest was reported as suitable harvest for maximum biomass yield [11,98]. Adler et al. [97] found 40% reduction in switchgrass biomass production when the harvest was delayed until spring. Other reports [24,99] were in line with findings of Adler et al. [97] where they found a 30% yield reduction from spring harvest. In contrast, Parrish and Fike [3] found no yield differences between November and February harvests in Virginia. Generally, biomass yield was reduced when harvest was delayed until after killing frost [24,47]; however, later harvest may ensure stand productivity and persistence of switchgrass. In north-central USA, harvesting after killing frost produced the highest yields [100]. In the same location Casler and Boe [101] found that a mid-August harvest reduced switchgrass stand density over time. In general, it is believed that switchgrass should not be harvested within 6 weeks of the first killing frost to ensure NSC translocation to the plant crowns for setting new tiller buds and maintaining stand productivity [47].

Cutting height is another important harvesting management practice that may influence final biomass yield [102]. Limited data is available on the influence of cutting height on the biomass production of switchgrass in the Northeast region of the United States. Existing reports suggest cutting heights between 15 to 25 cm will ensure switchgrass re-growth in the following year [103]. According to Henry et al. [104], the best switchgrass stand could be obtained from a cutting height of 23 cm in a single-cut system whereas in a two-harvest system, 8 cm would be the ideal harvesting height to gain maximum biomass yield. Several reports indicated that although cutting switchgrass low as 5-8 cm compared with 20-25 cm may result in higher biomass yield in the short term, biomass will be lowered in the following years due to intensified weed infestation [102,103,105]. Mitchell and Schmer [47] reported that cutting heights lower than 10 cm resulted in yield reductiondue to stand vigor loss. In a three year period Sadeghpour et al. [90] reported that cutting height of 7.5 cm out yielded cutting at 15 cm by 1 Mg ha-1 without increasing weed pressure.

Quality parameters of switchgrass as biofuel feedstock include energy content of grass, moisture, nutrients, and ash. Higher moisture and ash both reduce energy content, since higher moisture requires excess energy input to burn, and ash creates fouling in combustion equipment [106]. The presence of alkali metals and silicates in ash are major contributors to the production of slag, a thick black liquid material that forms when feedstock is burned at high temperatures. Slag coats the surfaces of machinery (furnaces, boilers, fluidized beds, etc.), causes fouling and prevents heat from being recovered [106,31], therefore making the burning process costly. Part of the appeal of switchgrass is that it can be used with existing technologies to supplement current energy production. It is imperative that the end product be used without causing high external costs to existing systems.

Harvesting management of switchgrass such as time of harvest may alter the concentration of unwanted nutrients present in the grass and therefore influence feedstock quality for combustion purpose. There is a general conformity in the literature that delaying the harvest of switchgrass until killing frost (after senescence), reduces N, phosphorus (P), potassium (K), ash, and other nutrients in the grass [91,107,108]. Lower ash content is associated with translocation of mobile nutrients from the above-ground tissue to the root structure [24]. It is reported that every 1% increase in ash concentration decreases the heating value by as much as 0.2 MJ kg ha-1 [109]. Nitrogen cycles down into the below-ground tissues at the end of the growing season [110]. This is due to the fact that switchgrass has evolved to go dormant at the onset of winter, translocates nutrients, including N, from above-ground tissues to the below-ground for re-growth in the succeeding season [3]. Adler et al. [97] found that delaying the harvest until spring resulted in higher energy content of the biomass because of moisture and ash content reduction. Direct baling of switchgrass requires moisture content of 15% or below [106]. In a multi-harvest study, Gorlitsky et al. [111] found 30% moisture reduction when harvest was delayed from midSeptember to mid-November; however, themoisture content from the delayed harvest was still high (29%) which makes it unsuitable for direct bailing. In another study, Sadeghpour et al. [90] concluded that delaying harvest until spring (midApril) can reduce moisture content to an acceptable level of 11 - 15%; however, this comes at the cost of a yield loss of about 25 to 30% which questions the suitability of harvesting in spring. A significant disparity of ash content of switchgrass across multiple locations ranging from 2.8 to 7.6% has been reported [106]. Adler et al. [97] showed that ash content reduced from 3.4 to 2.3% when the harvest was delayed until spring. Researchers [35,107] concluded that reduction in ash concentration from time of anthesis to killing frost harvest was related mainly due to greater proportion of grass stems at late season which contains less silica, a major component of ash, compared to leaves.

FERTILITY MANAGEMENT

Fertilization is perhaps the most unsettled aspect of switchgrass establishment and production [3]. Nitrogen fertilization is not recommended in the establishment year as it would encourage weed pressure and therefore not only increases establishment costs but also causes the economic risk associated with stand failure [67]. Sanderson and Reed [112] reported no biomass yield response to N application (22 and 112 kg ha-1) during the establishment year of “Alamo” switchgrass. They concluded that lack of switchgrass response to N fertilization was due to the ability of switchgrass to use available N in the soil. Reports have also indicated no significant response of switchgrass to P and K [67,113]. This is mainly due to the adequate levels of these elements in most agricultural soils. However, P and K fertilizers and lime are recommended to maintain soil nutrient balance during establishment and throughout production years [89].

Nitrogen

Nitrogen is a critical nutrient for production of biomass and typically the most limiting factor to plants productivity [114]. Managing N fertilizer application is important not only for optimum biomass production but also to maximize the NUE as well as feedstock quality. Excess N concentration in harvested switchgrass can be a liability by increasing the release of N oxide (NO and NO2 ) compounds into the atmosphere when combusted [3,114]. Most of studies on N management have been conducted on lowland switchgrass varieties in the Midwest, southern, and upper southeastern U.S.A. Nitrogen fertilizer recommendation are site specific and depend on weather, soil fertility level and management practices [67]. In a multi-location study throughout the upper southeastern USA, Lemus et al. [7] found that in a single-cut system, 50 kg N ha-1 would be sufficient for biomass production of switchgrass; however, a split application of N (100 kg N ha-1) is required in a 2-cut system to maintain grass productivity. It is reported that Alamo switchgrass yielded highest at N rate up to 224 kg ha-1. In a season of higher-thannormal rainfall, production was maximized at 168 kg N ha-1 [3]. Thomason et al. [94] found 448 kg N ha-1 application in a 3-cut system as the most suitable for maximum biomass production of Kanlow variety. However, multiple harvests each year resulted in a significant yield reduction in the succeeding years and they reported that a single harvest system over a four-year period at one of the locations of their study produced higher biomass compared with the 3-cut system with 448 kg N ha-1 fertilization. While yields were highest (18.0Mg ha-1) with 448 kg N ha-1 applied all in April and three harvests, no N application and harvesting three times produced almost as much total biomass (16.9Mg ha-1). This limited response to N is possibly explained by the evolution of switchgrass under low N conditions.

At the same location, Aravindhakshan et al. [115] reported that a single-cut system with only 69 kg N ha-1 was the most economical management practice for producing the greatest biomass production. Vogel et al. [98] tested N application rates up to 300 kg ha-1 for the Cave-in-Rock (a southern upland cultivar). They reported maximum yields at 120 kg N ha−1. Guertskyet al. [22] tested N up to 225 kg ha-1 at three harvest times (July, October, and December) and reported positive response of switchgrass biomass production to N fertilization. They found a 2-cut (July plus frost) harvest system the most productive however, higher N input was needed for this harvest system. In a recent multi-year-location study, Anderson et al. [116] recommended 56 kg N ha-1 in late fall to 112 kg N ha-1 in early spring to optimize switchgrass production. Harvesting switchgrass once a year after frost (December) has been suggested by several researchers [98,107,117]. In a study in Massachusetts on a 3-year old Cave-in-rock switchgrass Sadeghpour et al. [90] found that for a late-summer harvest (September) only a 67 kg N ha-1 was required to maintain stand productivity. No significant response of switchgrass yield to N fertilization in late-fall (November) and spring (April) harvests was detected. They concluded that perhaps less than 67 kg N ha-1 would be sufficient for growing high-yielding switchgrass in the state of Massachusetts. In another recent study, Pedroso et al. [118] found a linear response of switchgrass to N application where the greatest yields (9.7 and 13 Mg ha-1 yr-1) were obtained from the highest N fertilization rates (300 kg ha-1). They reported that the average NUE was between 30 to 44 kg biomass kg-1 N during 2009 and 2010 growing season. Sadeghpour et al. [90] found the average NUE to be from 14 up to 33% which was much lower than the averages reported by Bransby et al. [119]. Nitrogen-use efficiency can also be soil/site specific [3]. Lemus et al. [114] calculated different NUE for two different locations in Virginia. They reported that increasing the N rate at both sites could result in decreasing NUE at one site with no significant response in the other site. In a five-year experiment, Lemus et al. [120] in Iowa found 56 kg ha-1 an ideal N rate in terms of NUE. Overall, based on findings of Pedroso et al. [118], greater N fertilization would be required to sustain biomass production in warm ecoregions with greater yield potential.

Phosphorus, Potassium and pH

Limited research has been conducted on response of switchgrass to P and K fertilization [67]. Reports often suggested little [121] significant effect of these nutrients on switchgrass production which could be due to the inherent ability of switchgrass to use P that is available in the soil mainly through mycorrhizae symbiosis [122]. Mycorrhizae, by supplying the host plant with essential elements from the soil, can significantly increase plant growth [124]. Mycorrhize increase a plant’s ability to absorb water and growth limiting nutrients (notably P and N) through enhancing the root surface area in contact with the soil [124,125]. According to Brejda et al. [126] response of switchgrass to P and N was reduced when rhizospheremicroflora was back to stem-sterilized soils. In a recent study, Haque et al. [127] found no influence of P on switchgrass productivity and suggested a 135/0 kg N-P ha-1 application as the most economically viable fertilization system for switchgrass production. McKenna and Wolf [56] found small response of switchgrass to P fertilization when P levels in their soil test were low but only in the first year of their study.

Similar to P, switchgrass plants are efficient in their use of K [96]. Frequently little or no response of switchgrass to addition of K is reported [121,128]. In a greenhouse study, Friedrich et al. [129] found no yield improvement with applying K at rates up to 896 kg ha-1. In contrast, Kering et al. [84] reported that a combination application of 135 kg N and 68 kg K ha-1 produced the highest switchgrass biomass in Oklahama. They however, found no significant differences in biomass yield when comparingapplication of 68 kg K ha-1 alone with no fertilizer application.

There is a general conformity on tolerance of well-established switchgrass stands to many adverse environmental conditions including extreme pH. Reports on the influence of low pH on newly-established switchgrass seedlings are controversial. According to McLaughlin and Kszos [31] greenhouse studies in North Dakota showed a significant reduction in seedling survival in soil pH < 4.0 or > 8.0. Jung et al. [122] also reported 50% yield reduction on strong acidic (pH 4.3-4.9) soils compared with limetreated soils. In contrast to these findings, in other studies [130- 132] no limiting effect of soil acidity on switchgrass establishment has been found.

CONCLUSION

In the last 30 years, significant progress through dedicated research efforts has been made in developing switchgrass as a bioenergy crop. Although there is an improved understanding of the biology and agronomy of switchgrass, a few aspects of switchgrass establishment and production need further investigation. Reliable establishment methods and effective weed management practices to produce a harvestable biomass in the establishment year, appropriate nutrient management to enhance fertilizer efficiency, and biomass conversion methods are yet not fully determined. Best agronomic management practices coupled with genetics will result in high-yielding quality switchgrass for more efficient conversion.

REFERENCES

1. Colbran, N, Eide A. Biofuel, the environment, and food security: A global problem explored through a case study of Indonesia. Sustain Dev Law Pol. 2008; 9: 4-11.

2. Naik SN, Goud VV, Rout PK, Dalai AK. Production of first and second generation biofuels: a comprehensive review. Renew Sust Energ Rev. 2010; 14: 578-597.

3. Parrish DJ, Fike JH. The biology and agronomy of switchgrass for biofuels. Crit Rev Plant Sci. 2005; 24: 423-459.

4. Shastri YN, Hansen AC, Rodríguez LF, Ting KC. Switchgrass practical issues in developing a fuel crop. CAB Reviews 2012; 7: 1-14.

5. Balan V, Kumar S, Bals B, Chundawat S, Jin M, Dale B. Biochemical and thermochemical conversion of switchgrass to biofuels. Switchgrass, Green Energy and Technology. 2012; 153-185.

6. Gorlitsky LE. Management of switchgrass for the production of biofuel. Amherst (MA): University of Massachusetts, 2012.

7. Lemus R, Parrish DJ, Wolf DD. Nutrient uptake by “Alamo” switchgrass used as an energy crop. Bioenergy Res. 2009; 2: 37-50.

8. Ma Z, Wood CW, Bransby DI. Soil management impacts on soil carbon sequestration by switchgrass. Biomass and Bioenerg. 2000; 18: 469- 477.

9. Parrish DJ, Casler DM, Monti A. The evolution of switchgrass as an energy crop. In: Monti A, editor. Switchgrass a valuable biomass crop for energy. Springer-Verlag: London. 2012; 1-28.

10. Berti MT, Johnson BL. Switchgrass establishment as affected by seeding depth and soil type. Ind Crops Prod. 2013; 41: 289-293.

11. Sanderson MA, Read JC, Reed RL. Harvest management of switchgrass for biomass feedstock and forage production. Agron J. 1999; 91: 5-10.

12. Porter Jr CL. An anlaysis of variation between upland and lowland switchgrass Panicumvirgatum L. in central Oklahoma. Ecology. 1966; 47: 980-992.

13. Hultquist SJ, Vogel KP, Lee DJ, Arumuganathan K, Kaeppler S. Chloroplast DNA and nuclear DNA content variations among cultivars of switchgrass, Panicumvirgatum L. Crop Sci. 2013; 36: 1049-1052.

14. Forberg DB. Evaluation of seed vigor test for the establishment of switchgrass. Amherst (MA): University of Massachusetts, 2009.

15. Casler MD. Ecotypic variation among switchgrass populations from the nothern USA. Crop Sci. 2005; 45: 388-398.

16. Christian DG, Elbersen HW. Switchgrass (P. virgatum L.). Energy plant species. Their use and impact on environment and development, Vol. I., London: James and James. 1988; 257-263.

17. Elbersen HW, Ocumpaugh WR, Hussey MA, Sanderson MA, Tischler CR. Crown node elevation of switchgrass and kleingrass under low light. Crop Sci. 1998; 38: 712-716.

18. Jefferson PJ, McCaughey, WP. Switchgrass (Panicumvirgatum L.) cultivar adaptation, biomass production, and cellulose concentration as affected by latitude of origin. ISRN Agron. 2012; 2012: 1-9.

19. Van Esbroeck GA, Hussey MA, Sanderson MA. Variation between Alamo and Cave-In-Rock switchgrass in response to photoperiod extension. Crop Sci. 2003; 43: 639-643.

20. Wullschleger SD, Davis EB, Borsuk ME, Gunderson CA, Lynd LR. Biomass production in switchgrass across the United States: database description and determinants of yield. Agron. J. 2010; 102: 1158- 1168.

21. Sanderson MA, Adler PR, Boateng AA, Casler MD, Sarath G. Switchgrass as a biofuels feedstock in the USA. Can J Plant Sci. 2006; 86: 1315- 1325.

22. Guretzky JA, Biermacher JT, Cook BJ, Kering MK, Mosali J. Switchgrass for forage and bioenergy: Harvest and nitrogen rate effects on biomass yields and nutrient composition. Plant Soil. 2011; 339: 69-81.

23. Monti A, Venturi P, Elbersen HW. Evaluation of the establishment of lowland and upland switchgrass (Panicumvirgatum L.) varieties under different tillage and seedbed conditions in northern Italy. Soil Tillage Res. 2001; 63: 75-83.

24. Herbert SJ ,Gorlitsky L, Hashemi M, Sadeghpour A. Evaluating switchgrass varieties for biomass yield and quality in Massachusetts. Proceedings of the SunGrant National Conference: Science for Biomass Feedstock Production and Utilization, New Orleans, LA. 2012.

25. Blanco-Canqui H. Energy crops and their implications on soil and environment. Agron J. 2010; 102: 403-419.

26. Evers GW, Butler TJ. Switchgrass establishment on Coastal Plain soil. Proc American Forage and Grassland Council. 2000; 150-154.

27. Sadeghpour A, Hashemi M, DaCosta M, Weis SA, Jahanzad E, Herbert SJ. Cover crops, seeding methods and weed control for improving switchgrass establishment. ASA Northeastern regional meeting, Delaware, June 23-26; 2013.

28. Mitchell RB, Vogel KP, Sarath G. Managing and enhancing switchgrass as a bioenergy feedstock. Biofuels Bioprod Biorefin. 2008; 2: 530-539.

29. Mitchell RB, Vogel KP, Berdahl J, Masters RA. Herbicides for establishing switchgrass in the central and northern Great Plains. Bioenergy Res. 2010; 3: 321-327.

30. Perrin R, Vogel K, Schmer M, Mitchell R. Farm-scale production cost of switchgrass for biomass. Bioenergy Res. 2008; 1: 91-97.

31. McLaughlin SB, Kszos LA. Development of switchgrass (Panicumvirgatum) as a bioenergy feedstock in the United States. Biomass Bioenergy. 2005; 28: 515-535.

32. Madakadze IC, CoulmanBE, McElroy AR, Stewart KA, Smith DL. Evaluation of selected warm-season grasses for biomass production in areas with a short growing season. Bioresour.Technol. 1998; 65: 1-12.

33. Moser LE, Vogel KP. Switchgrass, Big Bluestem, and Indiangrass. In: Barnes RF, Miller DA, Nelson JC, editors. Forages Vol. 1, An Introduction to Grassland Agriculture, 5th edition. Ames, IA: Iowa State Univ. Press, 1995: 409-420.

34. Knapp AD Anderson BE, Moore KJ. An overview of seed dormancy in native warm-season grasses. In Native warm-season grasses: Research trends and issues. 2000; 107-122.

35. Mullen RE, Kassel PC, Bailey TB, Knapp AD. Seed dormancy and germination of switchgrass from different row spacings and nitrogen levels. J Applied Seed Prod. 1985; 3: 28-33.

36. Teel A, Barnhart S, Miller G. Management guide for the production of switchgrass for biomass fuel in southern Iowa. Ames, IA: Iowa State University Extension. 2003.

37. Zegada-Lizarazu W, Wullschleger SD, Nair SS, Monti A. Crop physiology. Monti A, editor. Switchgrass a valuable biomass crop for energy. Springer-Verlag: London. 2012; 55-86.

38. Zhang JH, Maun MA. Effects of partial removal of seed reserves on some aspects of seedling ecology of 7 dune species.Can. J. Botany.1991; 69: 1457-1462.

39. Sautter EH. Germination of switchgrass. J Range Manage. 1962; 15: 108-110.

40. Parrish DJ, Fike JH. Selecting, establishing, and managing switchgrass (Panicumvirgatum) for biofuels. Methods Mol Biol. 2009; 581: 27-40.

41. Shen ZX, Parrish DJ, Wolf DD, Welbaum GE. Stratification in switchgrass seeds is reversed and hastened by drying. Crop Sci. 2001; 41: 1546-1551.

42. Smart AJ, Moser LE. Morphological development of switchgrass as affected by planting date. Agron J. 1997; 89: 958-962.

43. Zarnstorff ME, Keys RD, Chamblee DS. Growth regulator and seed storage effects on switchgrass germination.Agron. J. 1994; 86: 667-672.

44. Shen ZX, Welbaum GE, Parrish DJ, Wolf DD. After-ripening and aging as influenced by anoxia in switchgrass (Panicumvirgatum L.) seeds stored at 60 deg C. Acta Hort. 1999; 504: 191-197.

45. AOSA. Rules for testing seed. J Seed Tech. 1993; 16: 1-113.

46. Gutormson TJ, Patin AL. Sources of laboratory test result variation in warm-season grasses. Seed Tech. 2002; 24: 52-61.

47. Mitchell R, Schmer M. Switchgrass harvest and storage. In: Monti A, editor. Switchgrass a valuable crop for energy. Springer-Verlag: London. 2012; 113-127.

48. Forberg DB, Herbert SJ, Prostack R, Hashemi M. Seed vigor test for the establishment of switchgrass. ASA, CSSA, and SSSA Annual meetings; 2009 Nov 1-5; Pittsburgh, PA.

49. Vogel KP, Masters RA. Frequency grid–a simple tool for measuring grassland establishment. J Range Manage. 2001; 54: 653-655.

50. Vogel, KP. Improving warm-season forage grasses using selection, breeding, and biotechnology. In: Anderson BE, Moore KJ, editors. Native Warm-Season Grasses: Research Trends and Issues, Madison, WI: CSSA. 2000; 83-106.

51. Schmer MR, Vogel KP, Mitchell RB, Moser LE, Eskridge KM, Perrin RK. Establishment stand thresholds for switchgrass grown as a bioenergy crop. Crop Sci. 2006; 46: 157-161.

52. Elbersen HW, Christian DG, El-Bassem N, Bacher W, Sauerbeck G, Alexopoulou E, et al. Switchgrass variety choice in Europe. Aspects of Appl Biol. 2001; 65: 21-28.

53. Oldfather S, Stubbendieck J, Waller SS. Evaluating revegetation practices for sandy cropland in the Nebraska sandhills. J Range Manage. 1989; 42: 257-259.

54. Potvin MA. Establishment of native grass seedlings along a topographic moisture gradient in the Nebraska sandhills. Amer Midland Naturalist.1993; 130: 248-261.

55. Wolf DD, Parrish DJ, Daniels WL, McKenna JR. Notill establishment of perennial, warm season grasses for biomass production. Biomass.1989; 20: 209-217.

56. McKenna JR, Wolf DD, Lentner M. No-till warm-season grass establishment as affected by atrazine and carbofuran. Agron. J. 1991; 83: 311-316.

57. Sadeghpour A, Hashemi M, Jahanzad E, Gorlitsky LE, Weis SA, Herbert SJ. Assessing various seedbed preparations on switchgrass establishment and production. ASA, CSSA, SSSA meetings, Tampa, FL, Nov 3-6; 2013.

58. Venturi P, Monti A, Elbersen HW. Soil tillage and switchgrass (P. virgatum L.) establishment. In: Proceedings of the Conference on Energy and Agriculture Towards the Third Millennium, Athens, June 2–5, 1999; 76-84.

59. Rehm GW. Importance of nitrogen and phosphorus for production of grasses established with no-till and conventional planting systems. J Prod Agric. 1990; 3: 333-336.

60. King MA, Waller SS, Moser LE, Stubbendieck JL. Seedbed effects on grass establishment on abandoned Nebraska Sandhills cropland. J Range Manage. 1989; 42: 183-187.

61. Harper CA, Morgan GD, Dixon CE. Establishing native warm-season grasses using conventional and no-till technology with various applications of Plateau herbicide.Proceedings of the 3rd Eastern Native Grass Symposium. Randall J, Burns JC, editors, Chapel Hill, NC: 2004; 63-70.

62. Hashemi M, Farsad A, Sadeghpour A, Weis SA, Herbert SJ. Cover crop seeding date influence on fall nitrogen recovery. J Plant Nutr Soil Sci. 2013; 176: 69-75.

63. Sadeghpour A, Jahanzad E, Esmaeili A, Hosseini MB, Hashemi M. Forage yield, quality and economic benefit of intercropped barley and annual medic in semi-arid conditions: Additive series. Field Crops Res. 2013; 148: 43-48.

64. Newman PR, Moser LE. Seedling root development and morphology of cool-season and warm-season forage grasses. Crop Sci. 1988; 28: 148-151.

65. Aiken GE, Springer TL. Seed size distribution, germination, and emergence of 6 switchgrass cultivars. J Range Manage. 1995; 48: 455- 458.

66. Evers GW, Parsons MJ. Soil type and moisture level influence on Alamo switchgrass emergence and seedling growth. Crop Sci. 2003; 43: 288- 294.

67. Sanderson MA, Schmer M, Owens V, Keyser P, Elbersen W. Crop management of switchgrass. In: Monti A, editor. Switchgrass a valuable biomass crop for energy. Springer-Verlag: London. 2012; 87-112.

68. Hsu FH, Nelson CJ. Planting date effects on seedling development of perennial warm-season forage grasses. I. Field emergence. Agron J. 1986; 78: 33-38.

69. Hsu FH, Nelson CJ. Planting date effects on seedling development of perennial warm-season forage grasses. II. Seedling growth. Agron J. 1986; 78: 38-42.

70. Hsu FH, Nelson CJ. Relationships between germination tests and field emergence of perennial warm-season forage grasses. In: Proceedings of XV Int. Grassl. Cong. Kyoto, Japan, 1985; 380-381.

71. Sanderson MA, Reed RL, McLaughlin SB, Wullschleger SD, Conger BV, Parrish DJ, et al. Switchgrass as a sustainable bioenergy crop. Bioresour Technol. 1996; 56: 83-93.

72. Buhler DD, Netzer DA, Riemenschneider DE, Hartzler RG. Weed management in short rotation poplar and herbaceous perennial crops grown for biofuel production. Biomass and Bioenergy. 1998; 14: 385- 394.

73. Boydston RA, Collins HP, Fransen SC. Response of three switchgrass (Panicumvirgatum) cultivars to mesotrione, quinclorac, and pendimethalin. Weed Technol. 2010; 24: 336-341.

74. Curran WS, Ryan MR, Myers MW, Adler PR. Effectiveness of sulfosulfuron and quinclorac for weed control during switchgrass establishment. Weed Technol. 2011; 25: 598-603.

75. Dawson JH, Rincker CM. Weeds in new seedlings of alfalfa (Medicago sativa) for seed production: Competition and control. Weed Sci. 1982; 30: 20-25.

76. Kelly AF. Principles of Seed Growing. In Seed Production of Agricultural Crops. New York: Wiley. 1988; 36-56.

77. Peters EJ, Linscott DL. Weeds and weed control. In: Hanson AA, Barns DK, Hill, RR, editors. Alfalfa and alfalfa improvement. ASA. Madison, WI, 1988; 705-735.

78. Putnam AR. Allelopathy: Problems and opportunities in Weed Management. In: Altieri MA, Liebman M, editors. Weed Managment in Agroecosystems: Ecological Approaches. Boca Raton, FL: CRC Press. 1988; 77-88.

79. Curran WS, Ryan MR, Myers MW, Adler PR. Effect of seeding date and weed control on switchgrass establishment. Weed Technol. 2012; 26: 248-255.

80. Bahler CC, Vogel KP, Moser LE. Atrazine tolerance in warm-season grass seedlings. Agron J. 1984; 76: 891-895.

81. Martin AR, Moomaw RS, Vogel KP. Warm-season grass establishment with atrazine.Agron. J. 1982; 74: 916-920.

82. Masters RA. Establishment of big bluestem and sand bluestem cultivars with metochlor and atrazine.Agron J. 1995; 87: 592-596.

83. Miesel JR, Renz MJ, Doll JE, Jackson RD. Effectiveness of weed management methods in establishment of switchgrass and a native species mixture for biofuels in Wisconsin. Biomass Bioenerg. 2012; 36: 121-131.

84. Kering MK, Huo C, Interrante SM, Hancock DW, Butler TJ. Effect of various herbicides on warm-season grass weeds and switchgrass establishment. Crop Sci. 2013; 53: 666-673.

85. Curran WS, Lingenfelter DD, Calvin DD, Tooker JF, Dillon JM. Forages pest management. In: Kirsten A, editor. Agronomy Guide. University Park, PA: Penn State University, College of Agricultural Sciences. 2008; 323-346

86. Rushing JB, Baldwin BS, Taylor AG, Owens VN, Fike JH, Moore KJ. Seed safening from herbicidal injury in switchgrass (Panicumvirgatum L.) establishment. Crop Sci. 2013; 53: 1650-1657.

87. Anderson WP. Weed Science: Principles and Applications. 3rd edn. Long Grove, IL: Waveland Press Inc. 1996.

88. Butler TJ, Kering MK, Huo C, Guretsky JA. Effect of safener, activated charcoal coated seed, and charcoal banding on establishment of switchgrass receiving pre-emergent herbicides. Forage Grazinglands. 2012.

89. Sokhansanj S, Mani S, Turhollow A, Kumar A, Bransby D, Lynd L, et al. Large-scale production, harvest and logistics of switchgrass (Panicumvirgatum L.) – current technology and envisioning a mature technology. Biofuels Bioprod Bioref. 2009; 3: 124-141.

90. Sadeghpour A, Gorlitsky LE, Hashemi M, Weis SA, Herbert SJ. Harvest management and nitrogen application rate influence on biomass yield, quality and nitrogen-use efficiency of switchgrass grown for combustion. ASA, CSSA, SSSA meetings, Tampa, FL, Nov 3-6; 2013.

91. Madakadze IC, Stewart KA, Peterson PR, Coulman BE, Smith DL. Cutting frequency and nitrogen fertilization effects on yield and nitrogen concentration of switchgrass in a short season area. Crop Sci. 1999; 39: 552-557.

92. Reynolds JH, Walker CL, Kirchner MJ. Nitrogen removal in switchgrass biomass under two harvest systems. Biomass Bioenergy. 2000; 19: 281-286.

93. Smart AJ, Moser LE, Vogel KP. Morphological characteristics of big bluestem and switchgrass plants divergently selected for seedling tiller number. Crop Sci. 2004; 44: 607-613.

94. Thomason WE, Raun WR, Johnson GV, Taliaferro CM, Freeman KW, Wynn KJ, et al. Switchgrass response to harvest frequency and time and rate of applied nitrogen. J Plant Nutr. 2005; 27: 1119-1226.

95. Fike JH, Parrish DJ, Wolf DD, Balasko JA, Green Jr JT, Rasnake M, et al. Long-term yield potential of switchgrass-for-biofuel systems. Biomass and Bioenergy. 2006; 30: 198 206.

96. Parrish DJ, Fike JH. Selecting, establishing, and managing switchgrass (Panicumvirgatum) for biofuels. Methods Mol Biol. 2009; 581: 27-40.

97. Adler PR, Sanderson MA, Boateng AA, Weimer PJ, Jung HG. Biomass yield and biofuel quality of switchgrass harvested in fall or spring. Agron J. 2006; 98: 1518-1525.

98. Vogel KP, Brejda JJ, Walters DT, Buxton DR. Switchgrass biomass production in the Midwest USA: Harvest and Nitrogen Management. Agron J. 2002; 94: 413-420.

99. Jannasch R., Quan Y, Samson R. A process and energy analysis of pelletizing switchgrass.Final report. Natural Resources Canada. 2001.

100. Mulkey VR, Owens VN, Lee DK. Management of switchgrass-dominated Conservation Reserve Program lands for biomass production in South Dakota. Crop Sci. 2006; 46: 712-720.

101. Casler MD, Boe AR. Cultivar x environment interactions in switchgrass. Crop Sci. 2003; 43: 2226-2233.

102. Trócsányi ZK, Fieldsend AF, Wolf DD. Yield and canopy characteristics of switchgrass (Panicumvirgatum L.) as influenced by cutting management. Biomass and Bioenergy. 2009; 33: 442-448.

103. Kiss Z, Fieldsend AF, Wolf DD. Yield of switchgrass (Panicumvirgatum L.) as influenced by cutting management. ActaAgronomica Hung. 2007; 55: 227-233.

104. Henry DS, Everett HW, Evans JK. Clipping effect on stand, yield and quality of three warm-season grasses. Proceedings of the Hill Lands Symposium Morgantown, WV: West Virginia University. 1976; 701- 704.

105. Anderson WP. Weed Science: Principles and Applications. 3rd edn. Long Grove, IL: Waveland Press Inc. 1996.

106. McLaughlin S, Bouton J, Bransby D, Conger B, Ocumpaugh W, Parrish D, et al. Developing switchgrass as a bioenergy crop. Pespectives on new crops and new uses. Alexandria, VA. ASHS Press.1996; 282-299.

107. Waramit N, Moore KJ, Heggenstaller AH. Composition of native warm-season grasses for bioenergy production in response to nitrogen fertilization rate and harvest date.Agron. J. 2011; 103: 655- 662.

108. Yang J, Worley E, Wang M, Lahner B, Salt DE, Saha M, et al. Natural variation for nutrient use and remobilization efficiencies in switchgrass. Bioenergy Res. 2009; 2: 257-266.

109. Cassida KA, Muir JP, Hussey MA, Read JC, Venuto BC, Ocumpaugh WR. Biofuel component concentration and yield of switchgrass in south central USA environments. Crop Sci. 2005; 45: 682-692.

110. Wilson DM, Dalluge LD, Rover M, Heaton EA, Brown RC. Crop management impacts biofuel quality: influence of switchgrass harvest time on yield, nitrogen and ash of fast pyrolysis products. Bioresour Technol. 2013; 6: 103-113.

111. Gorlitsky LE, Sadeghpour A, Hashemi M, Weis SA, Herbert SJ. Optimal fall harvest time for production of high-quality switchgrass grown for combustion. ASA, CSSA, SSSA Annual meetings; 2013 Nov 3-6; Tampa, FL.

112. Sanderson MA, Reed RL. Switchgrass growth and development: Water, nitrogen, and plant density effects. J Range Manage. 2000; 53: 221-227.

113. Parrish DJ, Fike JH. Selecting, establishing, and managing switchgrass (Panicumvirgatum) for biofuels. Methods Mol Biol. 2009; 581: 27- 40.

114. Lemus R, Parrish DJ, Abaye O. Nitrogen-use dynamics in switchgrass grown for biomass. Bioenergy Res. 2008a; 1: 153-162.

115. Aravindhakshan SC, Epplin FM, Taliaferro CM. Switchgrass, bermudagrass, flaccidgrass, and lovegrass biomass yield response to nitrogen for single and double harvest. Biomass and Bioenergy. 2011; 35: 308-319.

116. Anderson EK, Parrish AS, Voigt TB, Owens VN, Hong CH, Lee DK. Nitrogen fertility and harvest management of switchgrass for sustainable bioenergy feedstock production in Illinois. Ind Crops Prod. 2013; 48: 19-27.

117. Sanderson MA, Reed RL, McLaughlin SB, Wullschleger SD, Conger BV, Parrish DJ, et al. Switchgrass as a sustainable bioenergy crop. Bioresour Technol. 1996; 56: 83-93.

118. Pedroso GM, Hutmacher RB, Putnam D, Wright SD, Six J, van Kessel C, et al. Agron. J. 2013; 105: 311-320.

119. Bransby DI, McLaughlin SB, Parrish DJ. A review of carbon and nitrogen balances in switchgrass grown for energy. Biomass Bioenergy.1998; 14: 379-384.

120. Lemus R, Brummer EC, Burras CL, Moore KJ, Barker MF, Molstad NE, et al. Effects of nitrogen fertilization on biomass yield and quality in large fields of established switchgrass in southern Iowa. USA. Biomass Bioenerg. 2008b; 32: 1187-1194.

121. Brejda JJ. Fertilization of native warm-season grasses. In: Anderson BE, Moore KJ, editors. Native warm-season grasses: Research trends and issues. Native warm-season grasses: research trends and issues. Proceedings of the Native Warm-Season Grass Conference and Expo, Des Moines, IA, USA, 12-13 September 1996. 2000; 30: 177-200.

122. Jung GA, Shaffer JA, Stout WL. Switchgrass and big bluestem responses to amendments on strongly acid soil. Agron J. 1988; 80: 669-676.

123. Clark RB. Differences among mycorrhizal fungi for mineral uptake per root length of switchgrass grown in acidic soil. J Plant Nutr. 2002; 25: 1753-1772.

124. Clark RB, Zobel RW, Zeto SK. Effects of mycorrhizal fungus isolates on mineral acquisition by Panicumvirgatum in acidic soil. Mycorrhiza.1999; 9: 167-176.

125. Clark RB, Zobel RW, Zeto SK. Effects of mycorrhizal fungus isolates on mineral acquisition by Panicumvirgatum in acidic soil. Mycorrhiza.1999; 9: 167-176.

126. Brejda JJ, Moser LE, Vogel KP. Evaluation of switchgrassrhizospheremicroflora for enhancing seedling yield and nutrient uptake. Agron J. 1998; 90: 753-758.

127. Haque M, Biermacher JT, Kering MK, Guretzky JA. Economics of alternative fertilizer supply systems for switchgrass produced in phosphorus-deficient soils for bioenergy feedstock. BioEnergy Res. 2013; 6: 351-357.

128. Hall KE, George JR, Riedl RR. Herbage dry matter yields of switchgrass, big bluestem and indiangrass with N fertilization. Agron J. 1982; 74: 47-51.

129. Friedrich JW, Smith D, Schrader LE. Herbage yield and chemical composition of switchgrass as affected by N, S and K fertilization. Agron J. 1977; 69: 30-33.

130. Bona L, Belesky DP. Evaluation of switchgrass entries for acid soil tolerance. Commun Soil Sci Plant Anal 1992; 23: 1827-1841.

131. Harper J, Spooner AE. Establishment of selected herbaceous species on acid bauxite mine soils.Proceedings Symposium on Surface mining, hydrology, sedimentology and reclamation. Aluminum Company of America. Bauxite, AR: 1983; 413-422.

132. Hopkins AA, Taliaferro CM. Genetic variation within switchgrass populations for acid soil tolerance. Crop sci. 1997; 37: 1719-1722.

Received : 28 Aug 2013
Accepted : 11 Sep 2013
Published : 13 Sep 2013
Journals
Annals of Otolaryngology and Rhinology
ISSN : 2379-948X
Launched : 2014
JSM Schizophrenia
Launched : 2016
Journal of Nausea
Launched : 2020
JSM Internal Medicine
Launched : 2016
JSM Hepatitis
Launched : 2016
JSM Oro Facial Surgeries
ISSN : 2578-3211
Launched : 2016
Journal of Human Nutrition and Food Science
ISSN : 2333-6706
Launched : 2013
JSM Regenerative Medicine and Bioengineering
ISSN : 2379-0490
Launched : 2013
JSM Spine
ISSN : 2578-3181
Launched : 2016
Archives of Palliative Care
ISSN : 2573-1165
Launched : 2016
JSM Nutritional Disorders
ISSN : 2578-3203
Launched : 2017
Annals of Neurodegenerative Disorders
ISSN : 2476-2032
Launched : 2016
Journal of Fever
ISSN : 2641-7782
Launched : 2017
JSM Bone Marrow Research
ISSN : 2578-3351
Launched : 2016
JSM Mathematics and Statistics
ISSN : 2578-3173
Launched : 2014
Journal of Autoimmunity and Research
ISSN : 2573-1173
Launched : 2014
JSM Arthritis
ISSN : 2475-9155
Launched : 2016
JSM Head and Neck Cancer-Cases and Reviews
ISSN : 2573-1610
Launched : 2016
JSM General Surgery Cases and Images
ISSN : 2573-1564
Launched : 2016
JSM Anatomy and Physiology
ISSN : 2573-1262
Launched : 2016
JSM Dental Surgery
ISSN : 2573-1548
Launched : 2016
Annals of Emergency Surgery
ISSN : 2573-1017
Launched : 2016
Annals of Mens Health and Wellness
ISSN : 2641-7707
Launched : 2017
Journal of Preventive Medicine and Health Care
ISSN : 2576-0084
Launched : 2018
Journal of Chronic Diseases and Management
ISSN : 2573-1300
Launched : 2016
Annals of Vaccines and Immunization
ISSN : 2378-9379
Launched : 2014
JSM Heart Surgery Cases and Images
ISSN : 2578-3157
Launched : 2016
Annals of Reproductive Medicine and Treatment
ISSN : 2573-1092
Launched : 2016
JSM Brain Science
ISSN : 2573-1289
Launched : 2016
JSM Biomarkers
ISSN : 2578-3815
Launched : 2014
JSM Biology
ISSN : 2475-9392
Launched : 2016
Archives of Stem Cell and Research
ISSN : 2578-3580
Launched : 2014
Annals of Clinical and Medical Microbiology
ISSN : 2578-3629
Launched : 2014
JSM Pediatric Surgery
ISSN : 2578-3149
Launched : 2017
Journal of Memory Disorder and Rehabilitation
ISSN : 2578-319X
Launched : 2016
JSM Tropical Medicine and Research
ISSN : 2578-3165
Launched : 2016
JSM Head and Face Medicine
ISSN : 2578-3793
Launched : 2016
JSM Cardiothoracic Surgery
ISSN : 2573-1297
Launched : 2016
JSM Bone and Joint Diseases
ISSN : 2578-3351
Launched : 2017
JSM Bioavailability and Bioequivalence
ISSN : 2641-7812
Launched : 2017
JSM Atherosclerosis
ISSN : 2573-1270
Launched : 2016
Journal of Genitourinary Disorders
ISSN : 2641-7790
Launched : 2017
Journal of Fractures and Sprains
ISSN : 2578-3831
Launched : 2016
Journal of Autism and Epilepsy
ISSN : 2641-7774
Launched : 2016
Annals of Marine Biology and Research
ISSN : 2573-105X
Launched : 2014
JSM Health Education & Primary Health Care
ISSN : 2578-3777
Launched : 2016
JSM Communication Disorders
ISSN : 2578-3807
Launched : 2016
Annals of Musculoskeletal Disorders
ISSN : 2578-3599
Launched : 2016
Annals of Virology and Research
ISSN : 2573-1122
Launched : 2014
JSM Renal Medicine
ISSN : 2573-1637
Launched : 2016
Journal of Muscle Health
ISSN : 2578-3823
Launched : 2016
JSM Genetics and Genomics
ISSN : 2334-1823
Launched : 2013
JSM Anxiety and Depression
ISSN : 2475-9139
Launched : 2016
Clinical Journal of Heart Diseases
ISSN : 2641-7766
Launched : 2016
Annals of Medicinal Chemistry and Research
ISSN : 2378-9336
Launched : 2014
JSM Pain and Management
ISSN : 2578-3378
Launched : 2016
JSM Women's Health
ISSN : 2578-3696
Launched : 2016
Clinical Research in HIV or AIDS
ISSN : 2374-0094
Launched : 2013
Journal of Endocrinology, Diabetes and Obesity
ISSN : 2333-6692
Launched : 2013
Journal of Substance Abuse and Alcoholism
ISSN : 2373-9363
Launched : 2013
JSM Neurosurgery and Spine
ISSN : 2373-9479
Launched : 2013
Journal of Liver and Clinical Research
ISSN : 2379-0830
Launched : 2014
Journal of Drug Design and Research
ISSN : 2379-089X
Launched : 2014
JSM Clinical Oncology and Research
ISSN : 2373-938X
Launched : 2013
JSM Bioinformatics, Genomics and Proteomics
ISSN : 2576-1102
Launched : 2014
JSM Chemistry
ISSN : 2334-1831
Launched : 2013
Journal of Trauma and Care
ISSN : 2573-1246
Launched : 2014
JSM Surgical Oncology and Research
ISSN : 2578-3688
Launched : 2016
Annals of Food Processing and Preservation
ISSN : 2573-1033
Launched : 2016
Journal of Radiology and Radiation Therapy
ISSN : 2333-7095
Launched : 2013
JSM Physical Medicine and Rehabilitation
ISSN : 2578-3572
Launched : 2016
Annals of Clinical Pathology
ISSN : 2373-9282
Launched : 2013
Annals of Cardiovascular Diseases
ISSN : 2641-7731
Launched : 2016
Journal of Behavior
ISSN : 2576-0076
Launched : 2016
Annals of Clinical and Experimental Metabolism
ISSN : 2572-2492
Launched : 2016
Clinical Research in Infectious Diseases
ISSN : 2379-0636
Launched : 2013
JSM Microbiology
ISSN : 2333-6455
Launched : 2013
Journal of Urology and Research
ISSN : 2379-951X
Launched : 2014
Journal of Family Medicine and Community Health
ISSN : 2379-0547
Launched : 2013
Annals of Pregnancy and Care
ISSN : 2578-336X
Launched : 2017
JSM Cell and Developmental Biology
ISSN : 2379-061X
Launched : 2013
Annals of Aquaculture and Research
ISSN : 2379-0881
Launched : 2014
Clinical Research in Pulmonology
ISSN : 2333-6625
Launched : 2013
Journal of Immunology and Clinical Research
ISSN : 2333-6714
Launched : 2013
Annals of Forensic Research and Analysis
ISSN : 2378-9476
Launched : 2014
JSM Biochemistry and Molecular Biology
ISSN : 2333-7109
Launched : 2013
Annals of Breast Cancer Research
ISSN : 2641-7685
Launched : 2016
Annals of Gerontology and Geriatric Research
ISSN : 2378-9409
Launched : 2014
Journal of Sleep Medicine and Disorders
ISSN : 2379-0822
Launched : 2014
JSM Burns and Trauma
ISSN : 2475-9406
Launched : 2016
Chemical Engineering and Process Techniques
ISSN : 2333-6633
Launched : 2013
Annals of Clinical Cytology and Pathology
ISSN : 2475-9430
Launched : 2014
JSM Allergy and Asthma
ISSN : 2573-1254
Launched : 2016
Journal of Neurological Disorders and Stroke
ISSN : 2334-2307
Launched : 2013
Annals of Sports Medicine and Research
ISSN : 2379-0571
Launched : 2014
JSM Sexual Medicine
ISSN : 2578-3718
Launched : 2016
Annals of Vascular Medicine and Research
ISSN : 2378-9344
Launched : 2014
JSM Biotechnology and Biomedical Engineering
ISSN : 2333-7117
Launched : 2013
Journal of Hematology and Transfusion
ISSN : 2333-6684
Launched : 2013
JSM Environmental Science and Ecology
ISSN : 2333-7141
Launched : 2013
Journal of Cardiology and Clinical Research
ISSN : 2333-6676
Launched : 2013
JSM Nanotechnology and Nanomedicine
ISSN : 2334-1815
Launched : 2013
Journal of Ear, Nose and Throat Disorders
ISSN : 2475-9473
Launched : 2016
JSM Ophthalmology
ISSN : 2333-6447
Launched : 2013
Journal of Pharmacology and Clinical Toxicology
ISSN : 2333-7079
Launched : 2013
Annals of Psychiatry and Mental Health
ISSN : 2374-0124
Launched : 2013
Medical Journal of Obstetrics and Gynecology
ISSN : 2333-6439
Launched : 2013
Annals of Pediatrics and Child Health
ISSN : 2373-9312
Launched : 2013
JSM Clinical Pharmaceutics
ISSN : 2379-9498
Launched : 2014
JSM Foot and Ankle
ISSN : 2475-9112
Launched : 2016
JSM Alzheimer's Disease and Related Dementia
ISSN : 2378-9565
Launched : 2014
Journal of Addiction Medicine and Therapy
ISSN : 2333-665X
Launched : 2013
Journal of Veterinary Medicine and Research
ISSN : 2378-931X
Launched : 2013
Annals of Public Health and Research
ISSN : 2378-9328
Launched : 2014
Annals of Orthopedics and Rheumatology
ISSN : 2373-9290
Launched : 2013
Journal of Clinical Nephrology and Research
ISSN : 2379-0652
Launched : 2014
Annals of Community Medicine and Practice
ISSN : 2475-9465
Launched : 2014
Annals of Biometrics and Biostatistics
ISSN : 2374-0116
Launched : 2013
JSM Clinical Case Reports
ISSN : 2373-9819
Launched : 2013
Journal of Cancer Biology and Research
ISSN : 2373-9436
Launched : 2013
Journal of Surgery and Transplantation Science
ISSN : 2379-0911
Launched : 2013
Journal of Dermatology and Clinical Research
ISSN : 2373-9371
Launched : 2013
JSM Gastroenterology and Hepatology
ISSN : 2373-9487
Launched : 2013
Annals of Nursing and Practice
ISSN : 2379-9501
Launched : 2014
JSM Dentistry
ISSN : 2333-7133
Launched : 2013
Author Information X