首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 843 毫秒
1.
One-pass harvest equipment has been developed to collect corn (Zea mays L.) grain, stover, and cobs that can be used as bioenergy feedstock. Nutrients removed in these feedstocks have soil fertility implication and affect feedstock quality. The study objectives were to quantify nutrient concentrations and potential removal as a function of cutting height, plant organ, and physiological stage. Plant samples were collected in 10-cm increments at seven diverse geographic locations at two maturities and analyzed for multiple elements. At grain harvest, nutrient concentration averaged 5.5 g?N kg?1, 0.5 g?P kg?1, and 6.2 g?K kg?1 in cobs, 7.5 g?N kg?1, 1.2 g?P kg?1, and 8.7 g?K kg?1 in the above-ear stover fraction, and 6.4 g?N kg?1, 1.0 g?P kg?1, and 10.7 g?K kg?1 in the below-ear stover fraction (stover fractions exclude cobs). The average collective cost to replace N, P, and K was $11.66 Mg?1 for cobs, $17.59 Mg?1 for above-ear stover, and $18.11 Mg?1 for below-ear stover. If 3 Mg ha?1 of above-ear stover fraction plus 1 Mg of cobs are harvested, an average N, P, and K replacement cost was estimated at $64 ha?1. Collecting cobs or above-ear stover fraction may provide a higher quality feedstock while removing fewer nutrients compared to whole stover removal. This information will enable producers to balance soil fertility by adjusting fertilizer rates and to sustain soil quality by predicting C removal for different harvest scenarios. It also provides elemental information to the bioenergy industry.  相似文献   

2.
D. Kleinfeld  M.Y. Okamura  G. Feher 《BBA》1984,766(1):126-140
The electron-transfer reactions and thermodynamic equilibria involving the quinone acceptor complex in bacterial reaction centers from R. sphaeroides were investigated. The reactions are described by the scheme: We found that the charge recombination pathway of D+QAQ?B proceeds via the intermediate state D+Q?AQB, the direct pathway contributing less than approx. 5% to the observed recombination rate. The method used to obtain this result was based on a comparison of the kinetics predicted for the indirect pathway (given by the product kAD-times the fraction of reaction centers in the Q?AQB state) with the observed recombination rate, kobsD+ →D. The kinetic measurements were used to obtain the pH dependence (6.1 ? pH ? 11.7) of the free energy difference between the states Q?AQB and QAQ?B. At low pH (less than 9) QAQ?B is stabilized relative to Q?AQB by 67 meV, whereas at high pH Q?AQB is energetically favored. Both Q?A and Q?B associate with a proton, with pK values of 9.8 and 11.3, respectively. The stronger interaction of the proton with Q?B provides the driving force for the forward electron transfer.  相似文献   

3.
Alfalfa (Medicago sativa L.) is a promising bioenergy and bioproduct feedstock because of its high yield, N-fixation capacity, potential for planting in rotation with corn (Zea mays L.), and valuable protein co-product (leaf meal). Our objective was to examine the effect of growth environment on biomass yield, cellulosic ethanol traits, and paper pulp fiber characteristics of alfalfa stems. Landscape position (summit and mild slope), season of harvest (four harvests per season), and multiple years (2005 and 2006) provided environmental variation. Alfalfa stem samples were analyzed for cell wall carbohydrate and lignin concentration. Stems were subjected to dilute acid pre-treatment, enzymatic saccharification, and pulping processes to measure relevant cellulosic ethanol and paper production traits. Landscape position was not a significant source of variation for yield or any biomass quality trait. Yields varied among harvests in 2005 (1,410–3,265 kg ha?1) and 2006 (1,610–3,795 kg ha?1). All cell wall, conversion test, and paper production traits exhibited year by harvest interactions with no clear pattern. Total carbohydrates and lignin ranged from 440 to 531 g?kg?1 DM and from 113 to 161 g?kg-1 DM, respectively. Release of cell wall sugars by the conversion test ranged widely (419 to 962 g?kg?1 DM). Fiber traits were similarly variable with length and fine content ranging from 1.24 to 1.59 mm and from 15.2% to 21.9%, respectively. Utilizing alfalfa biomass for cellulosic ethanol and paper pulp production will involve dealing with significant feedstock quality variation due to growth environment.  相似文献   

4.
Crop residues like corn (Zea mays L.) stover perform important functions that promote soil health and provide ecosystem services that influence agricultural sustainability and global biogeochemical cycles. We evaluated the effect of corn stover removal from a no-till, corn-soybean (Glycine max (L.) Merr) rotation on soil greenhouse gas (GHG; CO2, N2O, CH4) fluxes, crop yields, and soil organic carbon (SOC) dynamics. We conducted a 4-year study using replicated field plots managed with two levels of corn stover removal (none; 55 % stover removal) for four complete crop cycles prior to initiation of ground surface gas flux measurements. Corn and soybean yields were not affected by stover removal with yields averaging 7.28 Mg ha?1 for corn and 2.64 Mg ha?1 for soybean. Corn stover removal treatment did not affect soil GHG fluxes from the corn phase; however, the treatment did significantly increase (107 %, P?=?0.037) N2O fluxes during the soybean phase. The plots were a net source of CH4 (~0.5 kg CH4-C ha?1 year?1 average of all treatments and crops) during the generally wet study duration. Soil organic carbon stocks increased in both treatments during the 4-year study (initiated following 8 years of stover removal), with significantly higher SOC accumulation in the control plots compared to plots with corn stover removal (0–15 cm, P?=?0.048). Non-CO2 greenhouse gas emissions (945 kg CO2-eq ha?1 year?1) were roughly half of SOC (0–30 cm) gains with corn stover removal (1.841 Mg CO2-eq ha?1 year?1) indicating that no-till practices greatly improve the viability of biennial corn stover harvesting under local soil-climatic conditions. Our results also show that repeated corn stover harvesting may increase N loss (as N2O) from fields and thereby contribute to GHG production and loss of potential plant nutrients.  相似文献   

5.
Definitions are proposed for the independent and joint contributions that the chemical groups A and B make to the free energy of association of the ligand A?B with a receptor. The definitions are independent of the choice of the standard state and are consistent with the basic thermodynamic cycle relating the association of the ligands A?B, A?Y and X?B to the receptor Rappaport 1976. The basic idea is the use of the excess free energy of association of the ligand A?Y over the free energy of association of the reference ligand X?Y as the measure of the “independent” contribution of the group A to the binding. This definition allows the free energy of association of the ligand A?B to be written as the sum of the independent contributions of the groups A and B, their joint contribution, and an invariant free energy of association of the reference ligand with any receptor. With the appropriate definition of the receptor-reference ligand complex, water can be chosen as the reference ligand. Using ΔG(A?OH)?AG(HOH), ΔG(H?B?H)?ΔG(HOH) and ΔG(HO?C)?ΔG(HOH) as the definitions of the “independent” contributions of the chemical groups A, B and C to the binding of the ligand A?B?C, the joint contribution of the groups A and C to the binding is ΔG(A?B?C) ? ΔG(A?B?H) ? ΔG(H-B-C) + ΔG(H?B?H).  相似文献   

6.
The temperature dependences of the P870+Q?A → P870QA and P870+Q?B → P870QB recombination reactions were measured in reaction centers from Rhodopseudomonas sphaeroides. The data indicate that the P870+Q?B state decays by thermal repopulation of the P870+Q?A state, followed by recombination. ΔG° for the P870+Q?A → P870+Q?B reaction is ?6.89 kJ · mol?1, while ΔH° = ?14.45 kJ · mol?1 and ?TΔS° = + 7.53 kJ · mol?1. The activation ethalpy, H3, for the P870+Q?A Δ P870+Q?B reaction is +56.9 kJ · mol?1, while the activation entropy is near zero. The results permit an estimate of the shape of the potential energy curve for the P870+Q?A → P870+Q?B electron transfer reaction.  相似文献   

7.
Harvesting crop residue needs to be managed to protect agroecosystem health and productivity. DAYCENT, a process-based modeling tool, may be suited to accommodate region-specific factors and provide regional predictions for a broad array of agroecosystem impacts associated with corn stover harvest. Grain yield, soil C, and N2O emission data collected at Corn Stover Regional Partnership experimental sites were used to test DAYCENT performance modeling the impacts of corn stover removal. DAYCENT estimations of stover yields were correlated and reasonably accurate (adjusted r 2?=?0.53, slope?=?1.18, p?<<?0.001, intercept?=?0.36, p?=?0.11). Measured and simulated average grain yields across sites did not differ as a function of residue removal, but the model tended to underestimate average measured grain yields. Modeled and measured soil organic carbon (SOC) change for all sites were correlated (adjusted r 2?=?0.54, p?<<?0.001), but DAYCENT overestimated SOC loss with conventional tillage. Simulated and measured SOC change did not vary by residue removal rate. DAYCENT simulated annual N2O flux more accurately at low rates (≤2-kg N2O-N ha?1 year?1) but underestimated when emission rates were >3-kg N2O-N ha?1 year?1. Overall, DAYCENT performed well at simulating stover yields and low N2O emission rates, reasonably well when simulating the effects of management practices on average grain yields and SOC change, and poorly when estimating high N2O emissions. These biases should be considered when DAYCENT is used as a decision support tool for recommending sustainable corn stover removal practices to advance bioenergy industry based on corn stover feedstock material.  相似文献   

8.
P. Gast  T. Swarthoff  F.C.R. Ebskamp  A.J. Hoff 《BBA》1983,722(1):163-175
The yield of the triplet state of the primary electron donor of Photosystem I of photosynthesis (PT-700) and the characteristic parameters (g value, line shape, saturation behavior) of the ESR signal of the photoaccumulated intermediary acceptor A have been measured for two types of Photosystem I subchloroplast particles: Triton particles (TSF 1, about 100 chlorophyll molecules per P-700) that contain the iron-sulfur acceptors FX, FB and FA, and lithium dodecyl sulfate (LDS) particles (about 40 chlorophyll molecules per P-700) that lack these iron-sulfur acceptors. The results are: (i) In Triton particles the yield of PT-700 upon illumination is independent of the redox state of A and of FX,B,A and is maximally about 5% of the active reaction centers at 5 K. The molecular sublevel decay rates are kx = 1100 s?1 ± 10%, ky = 1300 s?1 ± 10% and kz = 83 s?1 ± 20%. In LDS particles the triplet yield decreases linearly with concentration of reduced intermediary acceptors, the maximal yield being about 4% at 5 K assuming full P-700 activity. (ii) In Triton particles the acceptor complex A consists of two acceptors A0 and A1, with A0 preceding A1. In LDS particles at temperatures below ?30°C only A0 is photoactive. (iii) The spin-polarized ESR signal found in the time-resolved ESR experiments with Triton particles is attributed to a polarized P-700-A?1 spectrum. The decay kinetics are complex and are influenced by transient nutation effects, even at low microwave power. It is concluded that the lifetime at 5 K of P-700A0A?1 must exceed 5 ms. We conclude that PT-700 originates from charge recombination of P-700A?0, and that in Triton particles A0 and A1 are both photoaccumulated upon cooling at low redox potential in the light. Since the state P-700AF?X does not give rise to triplet formation the 5% triplet yield in Triton particles is probably due to centers with damaged electron transport.  相似文献   

9.
10.
This work aims to fully elucidate the effects of a trehalose glassy matrix on electron transfer reactions in cyanobacterial Photosystem I (PS I). Forward and backward electron transfer rates from A1A? and A1B? to FX, and charge recombination rates from A0?, A1B?, A1A?, FX?, and [FA/FB]? to P700+ were measured in P700–FA/FB complexes, P700–FX cores, and P700–A1 cores, both in liquid and in a trehalose glassy matrix at 11% humidity. By comparing CONTIN-resolved kinetic events over 6 orders of time in increasingly simplified versions of PS I at 480?nm, a wavelength that reports primarily A1A?/A1B? oxidation, and over 9 orders of time at 830?nm, a wavelength that reports P700+ reduction and A0? oxidation, assignments could be made for nearly all of the resolved kinetic phases. Trehalose-embedded PS I samples demonstrated partially arrested forward electron transfer. The fractions of complexes in which electron transfer did not proceed beyond A0, A1 and FX were 53%, 16% and 22%, respectively, with only 10% of electrons reaching the terminal FA/FB clusters. The ~10?μs and ~150?μs components in both liquid and trehalose-embedded PS I were assigned to recombination between A1B? and P700+ and between A1A? and P700+, respectively. The kinetics and amplitudes of these resolved kinetic phases in liquid and trehalose-embedded PS I samples could be well-fitted by a kinetic model that allowed us to calculate the asymmetrical contribution of the A1A? and A1B? quinones to the electrochromic signal at 480?nm. Possible reasons for these effects are discussed.  相似文献   

11.
Ryegrass, harvested before ear emergence, was ensiled in triplicate in laboratory silos with and without pre-ensiling treatments. These were: 4.7 M sulphuric acid (39 and 79 g kg?1 DM); formalin (28 and 58 g kg?1 DM); ‘Add-F’ (22.5 M formic acid; 36 and 71 g kg?1 DM); formalin (28 and 58 g kg?1 DM); ‘Farmline’, a commercial additive containing acids and formalin (31 g kg?1 DM); heat (30 and 60°C for 1 h); and Lactobacillus plantarum inoculum + glucose (4.6 g kg?1 DM). The ensilage period was 120 days. Samples of each silage were incubated, in artificial fibre bags, in the rumens of four Hereford-cross steers fitted with permanent rumen cannulae. Total nitrogen (TN) disappearance was measured after incubation periods of 1, 2, 4, 7, 12 and 24 h.Hydrochloric acid and the heating pre-treatments had no significant effect (P > 0.05) on the pH, water-soluble carbohydrates and fermentation acid values compared with those of the control. Fermentation acid production was suppressed to an increasing extent by formalin (28 g kg?1 DM), sulphuric acid (79 g kg?1 DM), formic acid (36 g kg?1 DM), formalin (58 g kg?1 DM) and formic acid (71 g kg?1 DM).The application of L. plantarum + glucose increased fermentation acids from 90.3 g kg?1 DM in the control to 227 g kg?1 DM, owing to a 10-fold increase in lactic acid content from 21.7 to 202 g kg?1 DM.All treatments, with the exception of 30°C heating, produced silages with significantly (P < 0.001) more total nitrogen as protein than the control. This ranged from 334 g kg?1 TN (60°C heating) to 748 g kg?1 TN (formalin; 58 g kg?1 DM).All treatments significantly (P < 0.05) reduced TN loss, compared with the control silage values, after a 7-h incubation period in the rumen. After 24 h, however, TN disappearance from the treated silages was not significantly different from that of the control with the exception of the two silages treated with formalin alone. With these silages, the proportion of TN disappearing was 55.0 and 40.0% for the low and high application rates, respectively, compared with a value of 74.0% for the control.  相似文献   

12.
In-field measurements of direct soil greenhouse gas (GHG) emissions provide critical data for quantifying the net energy efficiency and economic feasibility of crop residue-based bioenergy production systems. A major challenge to such assessments has been the paucity of field studies addressing the effects of crop residue removal and associated best practices for soil management (i.e., conservation tillage) on soil emissions of carbon dioxide (CO2), nitrous oxide (N2O), and methane (CH4). This regional survey summarizes soil GHG emissions from nine maize production systems evaluating different levels of corn stover removal under conventional or conservation tillage management across the US Corn Belt. Cumulative growing season soil emissions of CO2, N2O, and/or CH4 were measured for 2–5 years (2008–2012) at these various sites using a standardized static vented chamber technique as part of the USDA-ARS’s Resilient Economic Agricultural Practices (REAP) regional partnership. Cumulative soil GHG emissions during the growing season varied widely across sites, by management, and by year. Overall, corn stover removal decreased soil total CO2 and N2O emissions by -4 and -7 %, respectively, relative to no removal. No management treatments affected soil CH4 fluxes. When aggregated to total GHG emissions (Mg CO2?eq ha?1) across all sites and years, corn stover removal decreased growing season soil emissions by ?5?±?1 % (mean?±?se) and ranged from -36 % to 54 % (n?=?50). Lower GHG emissions in stover removal treatments were attributed to decreased C and N inputs into soils, as well as possible microclimatic differences associated with changes in soil cover. High levels of spatial and temporal variabilities in direct GHG emissions highlighted the importance of site-specific management and environmental conditions on the dynamics of GHG emissions from agricultural soils.  相似文献   

13.
Corn (Zea mays L.) stover was identified as an important feedstock for cellulosic bioenergy production because of the extensive area upon which the crop is already grown. This report summarizes 239 site-years of field research examining effects of zero, moderate, and high stover removal rates at 36 sites in seven different states. Grain and stover yields from all sites as well as N, P, and K removal from 28 sites are summarized for nine longitude and six latitude bands, two tillage practices (conventional vs no tillage), two stover-harvest methods (machine vs calculated), and two crop rotations {continuous corn (maize) vs corn/soybean [Glycine max (L.) Merr.]}. Mean grain yields ranged from 5.0 to 12.0 Mg ha?1 (80 to 192 bu ac?1). Harvesting an average of 3.9 or 7.2 Mg ha?1 (1.7 or 3.2 tons ac?1) of the corn stover resulted in a slight increase in grain yield at 57 and 51 % of the sites, respectively. Average no-till grain yields were significantly lower than with conventional tillage when stover was not harvested, but not when it was collected. Plant samples collected between physiological maturity and combine harvest showed that compared to not harvesting stover, N, P, and K removal was increased by 24, 2.7, and 31 kg ha?1, respectively, with moderate (3.9 Mg ha?1) harvest and by 47, 5.5, and 62 kg ha?1, respectively, with high (7.2 Mg ha?1) removal. This data will be useful for verifying simulation models and available corn stover feedstock projections, but is too variable for planning site-specific stover harvest.  相似文献   

14.
We have developed a relatively simple simultaneous saccharification and fermentation (SSF) technique to determine the ethanol production potential for large sets of biomass samples. The technique is based on soaking approximately 0.5 grams of a biomass sample in aqueous ammonia at room temperature and at atmospheric pressure for 24 h, then fermenting with Saccharomyces cerevisiae D5A for 24 h using Spezyme CP, for enzymatic hydrolysis of structural polysaccharides. We have tested the technique on a set of corn stover samples representing much of the genetic variability in the commercial corn hybrid population. The samples were weighed into modified Ankom filter bags (F57) before soaking to avoid biomass loss during the process. Fermentation samples were analyzed for ethanol after 24 h by HPLC. Percentages of theoretical maximum ethanol yields of the samples ranged between 44.9 and 73%. We observed that percentages of theoretical maximum ethanol yields were highly correlated (r 2?=?0.90) with acid detergent lignin concentration while a low correlation was observed between cellulose concentration and ethanol yield. Near infrared spectra of corn stover samples were also examined. The coefficient of determination (r 2) from regression of predicted versus measured percent theoretical maximum ethanol yield was 0.96. This result suggests that using NIRS is a promising method for predicting ethanol yield, but larger calibration sets are necessary for obtaining improved accuracy for larger sample populations. We conclude that the developed SSF technique could be applied to large numbers of biomass samples to rapidly estimate ethanol yields and to compare different biomass samples in terms of ethanol yields.  相似文献   

15.
The kinetic behaviour of intrinsic factor-vitamin B12 binding has been examined under varying conditions using an albuminised charcoal separation technique. The overall reaction obeys second order rate laws. The intrinsic factor considered alone obeys first order laws; the velocity of reaction of vitamin B12 is too fast for measurement by the technique described but by deduction obeys first order laws. Rate constants as three temperatures, (k2 at 25°C=1.56·108·mole?1·s?1) the activation energy (E=12.7 kJ·mole?1) and Arrhenius constant (A=2.7·1010 1·mole?1·s?1 have been calculated. There is the possibility of diffusion control of the reaction in which case the E and A values are invalid. The effect of pH on the reaction has been studied and the results discussed in relation to the pH studies of other workers whose results show disagreement. Albumin coated charcoal was shown to discriminate between intrinsic factor-vitamin B12 and free vitamn B12 over a wide pH range. The apparent under-estimation of intrinsic factor in dilute solution was shown to be due to adsorption of the intrinsic factor to plastic tubes.  相似文献   

16.
Rates of hemolysis of rabbit erythrocyte suspensions induced by P. parvum (prymnesin) have been measured colorimetrically at 25.5°C and pH 5.5. The data have been treated previously as consecutive first-order rate processes associated with the prolytic and lytic periods from which two specific rate constants have been obtained, k′ and , respectively. These constants have been related to those obtained by a computer-generated fit of the rate data (absorbance At, as a function of time t) with the rate equation Y = D[1 + exp((X ? B)C)] + E. Here Y equals At, X = time, t; D is equal to a spread factor, Ai ? A; C is the slope of the curve at the inflection point; B is the midpoint time value, i.e., the time at which At = D2; E is termed the off-set constant and is equal to A. Of these constants, B is directly related to the length of the prolytic period, and C?1 is directly related to the specific first-order rate constant for hemolysis, kψ.  相似文献   

17.
Abstract

A single-point substitution of the O4′ oxygen by a CH2 group at the sugar residue of A 6 (i.e. 2′-deoxyaristeromycin moiety) in a self-complementary DNA duplex, 5′- d(C1G2C3G4A5A6T7T8C9G10C11G12)2 ?3, has been shown to steer the fully Watson-Crick basepaired DNA duplex (1A), akin to the native counterpart, to a doubly A 6:T7 Hoogsteen basepaired (1B) B-type DNA duplex, resulting in a dynamic equilibrium of (1A)→←(1B): Keq = k1/k-1 = 0.56±0.08. The dynamic conversion of the fully Watson-Crick basepaired (1A) to the partly Hoogsteen basepaired (1B) structure is marginally kinetically and thermodynamically disfavoured [k1 (298K) = 3.9± 0.8 sec?1; δH°? = 164±14 kJ/mol;-TδS°? (298K) = ?92 kJ/mol giving a δG298°? of 72 kJ/mol. Ea (k1) = 167±14 kJ/mol] compared to the reverse conversion of the Hoogsteen (1B) to the Watson-Crick (1A) structure [k-1 (298K) = 7.0±0.6 sec-1, δH°? = 153±13 kJ/mol;-TδS°? (298K) = ?82 kJ/mol giving a δG298°? of 71 kJ/mol. Ea (k-1) = 155±13 kJ/mol]. A comparison of δG298°? of the forward (k1) and backward (k-1) conversions, (1A)→←(1B), shows that there is ca 1 kJ/mol preference for the Watson-Crick (1A) over the double Hoogsteen basepaired (1B) DNA duplex, thus giving an equilibrium ratio of almost 2:1 in favour of the fully Watson-Crick basepaired duplex. The chemical environments of the two interconverting DNA duplexes are very different as evident from their widely separated sets of chemical shifts connected by temperature-dependent exchange peaks in the NOESY and ROESY spectra. The fully Watson-Crick basepaired structure (1A) is based on a total of 127 intra, 97 inter and 17 cross-strand distance constraints per strand, whereas the double A 6:T7 Hoogsteen basepaired (1B) structure is based on 114 intra, 92 inter and 15 cross-strand distance constraints, giving an average of 22 and 20 NOE distance constraints per residue and strand, respectively. In addition, 55 NMR-derived backbone dihedral constraints per strand were used for both structures. The main effect of the Hoogsteen basepairs in (1B) on the overall structure is a narrowing of the minor groove and a corresponding widening of the major groove. The Hoogsteen basepairing at the central A 6:T7 basepairs in (1B) has enforced a syn conformation on the glycosyl torsion of the 2′- deoxyaristeromycin moiety, A 6, as a result of substitution of the endocyclic 4′-oxygen in the natural sugar with a methylene group in A 6. A comparison of the Watson-Crick basepaired duplex (1A) to the Hoogsteen basepaired duplex (1B) shows that only a few changes, mainly in α, σ and γ torsions, in the sugar-phosphate backbone seem to be necessary to accommodate the Hoogsteen basepair.  相似文献   

18.
Efficient methods for lignin characterisation are increasingly important as the field of lignin valorisation is growing with the increasing use of lignocellulosic feedstocks, such as wheat straw and corn stover, in biorefineries. In this study, we characterised a set of authentic lignin biorefinery samples in situ with no prior purification and minimal sample preparation. Lignin chemical formulas and lignin Fourier transform infrared (FTIR) spectra were extracted from mixed spectra by filtering out signals from residual carbohydrates and minerals. From estimations of C, H and O and adjustment for cellulose and hemicelluloses contents, the average chemical formula of lignin was found to be C9H10.2O3.4 with slight variations depending on the biomass feedstock and processing conditions (between C9H9.5O2.8 and C9H11.1O3.6). Extracted FTIR lignin spectra showed many of the same characteristic peaks as organosolv and kraft lignin used as benchmark samples. Some variations in the lignin spectra of biorefinery lignin residue samples were found depending on biomass feedstock (wheat straw, corn stover or poplar) and on pretreatment severity, especially in the absorbance of bands at 1267 and 1032 cm?1 relative to the strong band at ~1120 cm?1. The suggested method of FTIR spectral analysis with adjustment for cellulose and hemicellulose is proposed to provide a fast and efficient way of analysing lignin in genuine lignin samples resulting from biorefineries.  相似文献   

19.
Simulated experimental data were generated from error-free data following the equation y = A ? Be?k1 where A, B, and k are constants and were analyzed by iterative nonlinear regression using one of two basic published computer programs. The effect of the simulated experimental error in y on the precision of the computed constants A, B, and k was evaluated. The errors were either independent of y (simple errors) or proportional to y (relative errors) and outliers were sometimes introduced. Other factors investigated were the number of data points per regression, the range of values of y, and the effect of weighting the data. The results show that the errors in the computed constants, and particularly the rate constant k, may be considerably magnified with respect to the errors in the experimental data. The quantitative relationships that are presented are useful aids in the design of biochemical experiments in which the above equation is applicable.  相似文献   

20.
Quantifying actual and theoretical ethanol yields from biomass conversion processes such as simultanteous saccharification and fermentation (SSF) requires expensive, complex fermentation assays, and extensive compositional analyses of the biomass sample. Near-infrared reflectance spectroscopy (NIRS) is a non-destructive technology that can be used to obtain rapid, low-cost, high-throughput, and accurate estimates of agricultural product composition. In this study, broad-based NIRS calibrations were developed for switchgrass biomass that can be used to estimate over 20 components including cell wall and soluble sugars and also ethanol production and pentose sugars released as measured using a laboratory SSF procedure. With this information, an additional 13 complex feedstock traits can be determined including theoretical and actual ethanol yields from hexose fermentation. The NIRS calibrations were used to estimate feedstock composition and conversion information for biomass samples from a multi-year switchgrass (Panicum virgatum L.) biomass cultivar evaluation trial. There were significant differences among switchgrass strains for all biomass conversion and composition traits including actual ethanol yields, ETOHL (L Mg?1) and theoretical ethanol yields, ETOHTL (L Mg?1), based on cell wall and non-cell wall composition NIRS analyses. ETOHL means ranged from 98 to 115 L Mg?1 while ETOHTL means ranged from 203 to 222 L Mg?1. Because of differences in both biomass yields and conversion efficiency, there were significant differences among strains for both actual (2,534?C3,720 L ha?1) and theoretical (4,878?C7,888 L ha?1) ethanol production per hectare. It should be feasible to improve ethanol yields per hectare by improving both biomass yield and conversion efficiency by using NIRS analyses to quantify differences among cultivars and management practices.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号