+ model ARTICLE IN PRESS Marine Chemistry xx (2006) xxx – xxx www.elsevier.com/locate/marchem Intact protein modification and degradation in estuarine environments Lori C. Roth, H. Rodger Harvey * University of Maryland Center for Environmental Science, Chesapeake Biological Laboratory, Box 38, Solomons, MD 20688, USA Received in revised form 8 September 2005; accepted 11 October 2005 Abstract Proteins are the principal organic nitrogen-containing compounds of living biomass and by virtue of the readily cleaved amide bonds are believed to be very labile. Nevertheless, proteins have been found in multiple organic matter pools. Experimental incubations among three diverse estuarine environments (anoxic Chesapeake Bay, Lower Delaware Bay, and freshwater marsh) were used to examine the initial stages of protein hydrolysis and associated structural modification and degradation of the model protein bovine serum albumin (BSA). Size-exclusion chromatography combined with electrospray ionization mass spectrometry observed multiple proteinaceous products of lower molecular weight formed over the course of the incubation. Products ranged from 63 to 13 kDa, suggesting initial degradation via sequential hydrolysis. Complete hydrolysis of added protein occurred in all three environments within 48 h. Amino acid analysis of the intermediate products of degradation suggests that the initial process (hours) involves the selective removal of polar, charged amino acids by bacterial assemblages present. As degradation of the protein products continues, other functional groups are lost, leaving the overall amino acid composition of remaining material closely resembling that seen for the intact protein. D 2006 Elsevier B.V. All rights reserved. Keywords: Protein degradation; Amino acids; DOM; Carbon cycling; DON 1. Introduction Proteins are integral components of the dissolved nitrogen cycle. Because most nitrogen in plankton (believed to be the biggest source of DOM in marine waters) is present as protein (Lourenço et al., 1998), it constitutes a likely starting material for much of the organic nitrogen present in marine environments. The dominance of dissolved organic nitrogen in aquatic environments as the amide form (Knicker and Hatcher, * Corresponding author. Tel.: +1 410 326 7206; fax: +1 410 326 7341. E-mail address: harvey@cbl.umces.edu (H.R. Harvey). 1997; McCarthy et al., 1997) lends further support that proteins or their modified products are one source of this material. Although proteins were previously considered labile in marine environments and therefore not readily preserved (Hollibaugh and Azam, 1983), more recent evidence suggests that preservation of proteinaceous material in aquatic environments does occur (Nguyen and Harvey, 1997, 1998; Pantoja et al., 1997). This includes reports of intact proteins observed in ocean waters (Tanoue, 1995) and the slow rates of turnover seen for membrane proteins which might allow preservation (Nagata et al., 1998). The mechanisms suggested for protein preservation or recycling include abiotic processes such as condensation reactions which result 0304-4203/$ - see front matter D 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.marchem.2005.10.025 MARCHE-02312; No of Pages 13 ARTICLE IN PRESS 2 L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx in reduced bioavailability (Hedges, 1978; Nagata and Kirchman, 1997) and sorption to mineral surfaces which impede microbial attack (Hedges and Hare, 1987; Keil et al., 1994; Kirchman et al., 1989; Lee and Ruckenstein, 1988). Additional mechanisms, including hydrophobic interactions (Nguyen and Harvey, 2001) and encapsulation (Knicker and Hatcher, 1997), allow for the stabilization of proteinaceous material over longer timescales. In all these processes, however, there remains a poor understanding of the first steps in protein recycling, the products produced, and their longevity in the environment. It is a matter of debate whether intermediates created during bacterial degradation of protein can be found in marine environments. Hollibaugh and Azam (1983) did not find intermediate peptides during the degradation of BSA; however, Keil and Kirchman (1991a) and Pantoja and Lee (1999) did observe what they considered intermediates during protein degradation based on the presence of high molecular weight amino acids. In addition, dissolved proteins have been observed in a range of natural waters (Nguyen and Harvey, 1998; Tanoue, 1995; Tanoue et al., 1996), suggesting that immediate hydrolysis is not universal. The major focus of this study is the examination of the degradation and potential modification of protein mediated by microbial processes. We investigated the degradation mechanism of dissolved protein and modification as evidenced by structural changes in a model protein in the presence of natural microbial communities. Experimental incubations were used to provide a comparison of the protein degradation and structural characteristics of intermediate peptide products during the degradation process. Molecular weight changes of intact products were followed over time while amino acid analysis was used to determine if alterations in molecular weight were accompanied by changes in major amino acid distribution. The goal was to examine the initial process of protein degradation by natural microbial consortia and formation of early degradation products which might contribute to the dissolved organic nitrogen pool in aquatic systems. 2. Materials and methods 2.1. Protein degradation incubations Field sampling took place from August 27 to 31, 2002, at two locations in the Delaware Bay system characterized by differing physiochemical characteris- tics. The Marsh site is a freshwater marsh (2) and is heavily influenced by tidal cycles with significant amounts of humic and organic material. The Delaware Bay site is near the mouth of the estuary and characterized by high salinity (31.5) and high productivity (Sharp et al., 1982; Pennock and Sharp, 1986). A third site included the deep channel of the nearby Chesapeake Bay which encounters seasonal anoxia, with a salinity of 15. Anoxic waters were present within 3 m of the surface during the time of sampling (G. Luther, unpublished results) and maintained under anoxic conditions throughout the incubation period. Additional incubations used water collected from the lower Patuxent River (salinity 10.5) near the entrance to the mesohaline portion of the Chesapeake Bay. At each site, one 20 L carboy was used to collect water via pumping from 1 m depth (or 10 m in anoxic waters) which was filtered through a 0.2 Am polycarbonate filter to remove all particles including phytoplankton and bacteria. A 10% (v/v) inoculum of water prefiltered through a 3 Am filter containing the natural bacterial community was added to the 0.2 Am filtered water. Bovine serum albumin (15 nM L 1) was added to the carboy containing the natural bacterial assemblage. The experimental incubations were sampled at 12, 24, 36, and 48 h, with aliquots removed and frozen at 70 8C until concentration and analysis. An aliquot of the 0.1 Am filtered surface water from the Patuxent River was used as an abiotic control to investigate the potential for short-term protein modification and was sampled at 0 and 92 h. From each sample, 60 mL was removed and processed as described below. For parallel experiments, Patuxent River water was collected at 0, 6, 12, 24, 36, 48, and 92 h and then processed as described below. 2.2. Protein isolation Dissolved protein in incubation and control waters was isolated and concentrated using ultrafiltration with centrifugal filters containing a nominal 5 kDa membrane (Millipore). Each aliquot was initially filtered through a 0.45 Am PDVP membrane filter (Gelman Sciences) followed by concentration. The samples were centrifuged (15,000g) at room temperature to reduce the volume to 150 AL. The retained concentrate was then rinsed with 2 mL of 0.16% sodium deoxycholate to minimize absorption to the filter membrane followed by two rinses of 2 mL Nanopure water to remove excess deoxycholate. The samples was transferred using 350 AL of 50 mM NH4CO3 ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx buffer (pH 7.0) amended with guanidined HCl (3 M final concentration) to minimize protein aggregation and stored at 70 8C until analysis. 2.3. Quantification and separation of proteins and peptides Size-exclusion chromatography was performed at room temperature under native conditions, (i.e. mobile phase without a denaturing agent) using a Superdex 200 HR column (30 cm 10 mm, Amersham Pharmacia) operating with 50 mM NH4HCO3 mobile phase at 0.5 mL min 1. Subsamples were injected and elution was monitored by fluorescence (k ex = 280 nm; k em = 340 nm). Peaks of interest were collected by fraction collection and stored at 70 8C until prepared for further analysis. The performance of the column and molecular weight calibration was constructed using a set of known molecular weight proteins including aprotinin (65 kDa), cytochrome c (12.4 kDa), carbonic anhydrase (29 kDa), bovine serum albumin (67 kDa), and blue dextran (2000 kDa-void volume) and a premixed gel permeation chromatography standard including thyroglobulin (670 kDa), gamma globulin (158 kDa), ovalbumin (44 kDa), myoglobin (17 kDa), and vitamin B12 (1.35 kDa). A calibration curve of these molecular weight standards was used to estimate molecular weights of unknowns based on retention time and repeated after each sample suite. To quantify BSA and its potential products, known concentrations of BSA were injected in equal volumes to develop a calibration curve used to estimate protein concentrations observed (Fig. 1). Molecular weight resolution of the column and flow rates used in the present study limited fraction collection to a minimum of 0.5 minute peak widths for amino acid analysis, equivalent to ~25 kDa based on protein calibrants. BSA concentration (µM) 0.60 y = 1.42x + 3.65 0.45 R2= 0.965 0.30 0.15 0 0 5 10 Peak Area 15 20 25 (x103) Fig. 1. Calibration curve for pure BSA verses fluorescence peak area used to determine unknown protein concentrations. 3 2.4. Molecular weight analysis by ESI-LC-MS Mass determination measurements were performed using the Agilent 1100 ion trap mass spectrometer with an electrospray ionization interface. Concentrated protein samples (752 nM) were infused at 400 AL min 1, using a mobile phase of 50 mM NH4HCO3 with 0.1% formic acid. The instrument was first tuned to get the optimal MS conditions for protein responses. The MS conditions used for all protein infusions are as follows: capillary voltage, 150–180 V; trap drive, 75–80 V; dry gas temperature, 350 8C; nebulizer gas pressure, 15.0 psi. The mass spectra were registered in normal scan mode (m / z, 200– 2000; scan time, 5 scans s 1, maximum accumulation time, 50 ms, ICC target, 30,000). As the protein was being infused, the instrument was tuned for final optimization parameters to obtain the highest responses for each protein product examined. Because proteins are multiple-charged compounds, they acquire a range of charge states yielding a series of masses for an individual compound. Deconvolution software (Agilent technologies) allowed statistical transform of the envelope of multiple-charged products into a singly charged parent spectrum for molecular weight determination. 2.5. Amino acid analysis To provide additional information on amino acid composition of observed degradation products, total amino acid analysis was conducted on isolated products using methods previously described (Mannino and Harvey, 2000. Briefly, individual samples were transferred to a 4 mL amber vial, a known amount (68.9 AM) of g-methylleucine added as an internal standard, and then dried with nitrogen gas. The samples were then redissolved in 0.5 mL of sequanal grade HCl (Sigma Chemical Co), capped under nitrogen, incubated at 150 8C for 2 h. Acid was removed from hydrolyzed samples under a gentle stream of nitrogen at 45 8C to approach dryness. The samples were subsequently derivatized to trifluoracyl isopropyl esters and analyzed by gas chromatography and gas chromatography/mass spectrometry (GC/MS) as described by (Silfer et al., 1991). Amino acid concentrations were determined based on the internal standard and an external standard commercial mix of the amino acids (Sigma) used for determination of yields. Variable responses in individual amino acids were corrected by comparison of peak areas of each amino acid to g-methylleucine and calcula- ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx Protein concentration (µg/ml) 4 Delaware Bay that sample processing did not alter amino acid composition. The average (n = 5) coefficient of variation for individual amino acids between these two approaches was F2.3%. Chesapeake Bay Marsh 1 0.5 3. Results 0 0 20 40 3.1. Comparison of dissolved protein hydrolysis among environments 60 Time (hrs) Fig. 2. Time course of degradation for dissolved protein determined from ultrafiltered (N5 kDa) concentrates in three different natural environments. The anoxic Chesapeake Bay, lower Delaware Bay and freshwater marsh incubations all showed a rapid decrease in BSA concentration from 0 to 24 h, with most rapid loss seen in freshwater marsh waters (Fig. 2). By 24 h, BSA was not detected in unconcentrated marsh incubations and present only in trace amounts at tion of recoveries. Replicate BSA samples were also dissolved and concentrated using ultrafiltration and compared to direct analysis of intact BSA to ensure Chesapeake Bay Delaware Bay A Marsh C B Fluorescence (λex = 280 nm; λem= 340 nm) 12 hr 24 hr 36 hr 48 hr 10 20 30 40 Elution time (min) All 0 hr 10 10 20 30 40 Elution time (min) 10 20 30 40 Elution time (min) Abiotic 92 hours. 20 30 40 Elution time (min) 10 20 30 40 Elution time (min) Fig. 3. Size-exclusion chromatograms of concentrated dissolved proteins and degradation products present during BSA degradation in the three different aquatic environments: a) Chesapeake Bay b) lower Delaware Bay and c) Freshwater Marsh. Abiotic controls after 92 h are shown as a separate inset. All three sites showed identical distributions of protein at time 0 with one example shown as a separate inset for illustration. ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx 3.2. ESI MS analysis An additional incubation using Patuxent River estuarine water was used to examine the potential for more rapid changes than seen by 24 h for all three sites and provide adequate material for detailed analysis. Incubation procedures were identical, but larger sample volumes allowed increased sample amounts required for preparative SEC followed by ESI MS analysis. Mass determination for concentrated protein products was performed following size exclusion chromatography (SEC). The addition of BSA to incubation waters resulted in a rapid appearance of a lower molecular weight product, seen as a mixture of BSA (66.2 kDa) plus a 64.8 kDa product within minutes of addition (Fig. 4). This mass shift corresponds to an average loss of approximately three amino acids from BSA. Within 6 h, a subset of lower molecular weight products (average mass of 63.3 kDa) was also observed, corresponding to a loss of approximately eleven amino acids. Intermediate products were in the same range with slight shifts in molecular weights at 12 h. By 12 h an additional smaller intermediate product was observed at 45.1 kDa ( 142 amino acids). There were also shifts in molecular weights from BSA (66.3, 1 amino acid and 57.7 kDa, 58 amino acids). After 24 hours, the largest peak observed had a molecular weight of 57.9 kDa ( 57 amino acids). Interestingly, an additional hydrolysis product was seen at 12 h (45.1 kDa) and observed through 24 h with similar intensity. The 0 hr 66,194 Da 64,802 Da (-11aa) 6 hr 66,379 Da 27.287 Relative Abundance 36 h in the Chesapeake Bay and lower Delaware Bay incubations. Despite the rapid recycling of dissolved protein by the microbial community, an examination of the molecular weight distributions of sample concentrates found a variable pattern of initial products formed among the three environments (Fig. 3). In all three environments, both higher and lower molecular weight products (relative to BSA) were seen by 12 h, with distributions dependent upon time and site. Incubations in anoxic Chesapeake Bay waters (Fig. 3a) showed a consistent loss throughout the incubation period, with a rapid loss of BSA accompanied by the appearance of multiple products through 24 h. By 36 h most were lost, and only small amounts of high molecular weight products remained at 48 h. In the lower Delaware Bay site (Fig. 3b), modified products were observed at 12 h ranging from 10 to N800 kDa in size. The distribution of modified products changed little at 24 h, with a substantial decrease in proteinaceous material by 36 h. Only the largest molecular weight masses which might represent aggregated material remained in the incubation at 48 h. In contrast, the freshwater marsh incubation differed significantly from the other sites (Fig. 3c) in that high molecular weight aggregates or modified products dominated the degradation products seen. As with incubations at other sites, a number of products were observed at 12 h ranging from 17 to 500 kDa with high molecular weight material dominant. While most products were subsequently lost from marsh waters by 24 h, a high molecular weight (N 1000 kDa) fraction remained through 36 h. Unlike either Chesapeake or Delaware incubations, marsh incubations showed little evidence of remaining protein or its modified products at 48 h. In the control (i.e. abiotic) incubation, BSA was the dominant peak with only small amounts of high molecular material (likely BSA dimers) also seen (Fig. 3, 92 h abiotic inset), confirming that biotic processes were responsible for the rapid changes seen in the bacterized incubations. 5 12 hr 63,339 Da (-20aa) 57,650 Da (-58aa) 45,117 Da (-142aa) 24 hr 57,893 (-57aa) 45,117 Da (-142aa) 10 20 30 40 Elution time (min) Fig. 4. Size exclusion chromatography of dissolved protein (BSA) degradation in Patuxent River waters. The shaded boxes indicate the peaks collected and further analyzed by ESI MS. The dominant mass of the peak is shown (in Da) and calculated by deconvolution. The second value estimates the number of amino acids lost from the original protein to achieve the calculated mass. Chromatograms are not to the same scale to allow degradation products to be viewed. ARTICLE IN PRESS 6 L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx longevity of this product in the incubation while others showed rapid loss indicates that its further degradation may be at a slower rate than its initial formation. 3.3. Amino acid analysis Previous work has observed that total amino acid composition generally remains invariant during organic matter degradation (Henrichs and Farrington, 1987; Nguyen and Harvey, 1994a, 1997; Siezen and Mague, 1978). To examine if this holds true during the initial stages of protein hydrolysis, the major amino acid composition of the total high molecular weight dissolved fraction (N 5 kDa–0.45 Am) amended by BSA and isolated products observed during short terms exposure to natural microbial assemblages were compared. 3.3.1. Total amino acids in dissolved protein Among individual amino acids, the modified products seen in sample concentrates showed only small shifts in composition over time (Table 1). To more easily compare functional changes in the products observed, individual amino acids were grouped by side chain functionality and summed. Using this approach, the model protein BSA contains 42% charged, polar amino acids, 51% nonpolar amino acids, and 7% uncharged, polar amino acids (Fig. 5). Within 1 h of the addition of BSA to the marsh waters, the distribution of major amino acids comprised charged, polar amino acids (68%), followed by nonpolar amino acids (30%) and uncharged, polar amino acids (2%) (Fig. 5-top panel). By 48 h when the incubation was halted, the amino acid distribution had shifted to 71% nonpolar, 24% charged polar amino acids and 5% uncharged polar amino acids. For the overall incubation, there was an increase among major nonpolar amino acids and a decrease in those with polar, charged side chains; uncharged, polar amino acids remained unchanged. Among individual nonpolar amino acids, glycine and alanine increased significantly during the incubation, while valine, phenylalanine, leucine, proline, and isoleucine remained unchanged. Although leucine did not change in molar contribution through the incubation, it was lower than that seen in BSA present in abiotic controls. The decrease in amino acids with the polar charged side chain appeared mainly due to a large decrease in aspartic acid; lysine remained largely unchanged, and glutamic acid was variable in amount throughout the incubation (Fig. 5). The Chesapeake Bay incubation showed similar trends in amino acid distribution to those seen in the marsh incubation (Fig. 5—middle panel). At 12 h the bulk of the amino acids was made up by the charged, polar amino acids (64%), followed by nonpolar amino acids (33%) and uncharged, polar amino acids (3%). At the end of the experiment (24 h) the amino acid distribution shifted to 66%, 30%, and 4%, for nonpolar amino acids, charged, polar amino acids, and uncharged polar amino acids, respectively. Again there was a small increase in nonpolar amino acids, a decrease in amino acids with polar, charged side chains. For the nonpolar amino acids, glycine and alanine increased significantly during the incubation, while phenylalanine, leucine, valine, proline, and isoleucine remained unchanged (Table 1). All amino acids with polar charged side chains decreased through the incubation (Fig. 5). By contrast with freshwater or anoxic waters, amino acid composition in the lower Delaware Bay incubation followed a markedly different pattern (Fig. 5—lower Table 1 Major amino acid composition (as mole percent) of total dissolved protein (5 kDa–0.45 um) fractions collected and concentrated at various time after addition of BSA to Marsh, Chesapeake Bay, and Lower Delaware Bay waters Marsh Chesapeake Bay Lower Delaware Bay Collection time (h) 12 24 36 48 12 24 36 48 12 24 36 48 THR PRO GLY ALA VAL LEU ILE PHE ASP + ASN GLU + GLN LYS 1.51 2.15 3.46 9.68 4.57 2.81 2.51 4.71 4.15 64.45 0.00 8.37 3.69 22.75 10.78 5.76 8.01 1.81 4.94 28.43 4.23 1.23 5.37 7.45 2.28 2.20 3.26 1.01 3.60 6.82 26.33 24.11 17.58 5.18 3.53 32.04 20.28 4.03 5.46 1.35 4.00 17.95 4.69 1.47 3.14 6.29 3.08 3.82 6.20 9.86 1.90 2.06 45.21 10.82 7.63 5.81 4.43 7.15 18.89 3.03 14.25 1.20 8.72 27.99 6.43 2.10 0.03 4.78 11.76 11.26 1.64 4.77 3.20 2.41 16.88 32.77 10.49 3.92 17.94 12.10 13.93 5.55 10.01 3.31 2.80 24.01 4.64 1.81 1.91 3.75 4.14 24.02 6.14 13.28 1.78 9.49 23.88 10.73 0.87 2.09 4.40 7.75 10.89 5.56 10.70 1.51 7.92 34.64 11.94 2.59 3.96 3.46 17.67 16.73 2.85 3.10 1.38 6.00 29.67 8.20 6.98 3.78 3.58 29.50 12.17 5.55 9.63 2.56 6.91 17.48 6.69 2.15 Trp is lost during acid hydrolysis, Met and His were highly variable and excluded. ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx 7 100 Uncharged, Polar Non polar Charged, Polar Mole Percent 80 60 40 20 0 BSA 12 24 36 48 BSA 12 24 36 48 Abiotic 80 60 40 20 0 Abiotic 100 Mole Percent 80 60 40 20 0 BSA 12 24 36 48 Abiotic Fig. 5. The amino acid distribution of dissolved protein (defined as N5 kDa) present in Marsh (top panel), Chesapeake Bay (middle panel) and lower Delaware Bay (bottom panel) waters at various time points over the 48-h incubation time. Amino acids are grouped into major functional groups to illustrate the changing character of the total protein pool. panel). The major amino acids after 12 h were the nonpolar amino acids (63%), followed by the amino acids with polar charged side chains (35%), and uncharged, polar amino acids (2%). This composition showed only minor shifts over the incubation period with a distribution of 70%, 26%, and 4%, respectively. Among the amino acids with nonpolar side chains, glycine showed a significant increase while leucine and alanine showed a significant decrease in the incubation, and valine, isoleucine, proline, and phenylala- nine remained largely unchanged. Glycine and alanine were relatively higher by 48 h than BSA with leucine reduced. For the polar, charged amino acids, aspartic acid showed a significant decrease over time, while lysine increased over the incubation and unlike the other incubations, glutamic acid remained relatively unchanged. Although lysine appeared to increase over the incubation period, it remained lower than the BSA standard at all time points. The overall amino acid distribution of total hydrolysable amino acids within ARTICLE IN PRESS 8 L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx the 5 kDa to 0.45 Am range shows a decrease in charged, polar amino acids. 3.3.2. Amino acid composition of modified products Identified products of protein degradation seen in SEC were collected and analyzed individually. For the marsh incubations, size-exclusion peaks were collected based on molecular weight fractions N BSA (100–1000 kDa), b BSA (2–52 kDa), and compared to BSA in major amino acid composition (Table 2). To compare changes in overall character, the amino acids were again grouped by side chain functionality (Fig. 6). In the marsh incubation at 12 h (Table 2), the dominant group of THAA in the NBSA fraction were amino acids with the charged, polar side chains (51%), followed by nonpolar amino acids (43%), and then the uncharged, polar amino acids (6%). This distribution is very similar to that seen for BSA (51%, 42%, and 7%, charged polar, nonpolar, uncharged polar, respectively). By contrast, modified products with molecular weights less than BSA at 12 and 24 h were very different from BSA, with amino acid distribution of 16%, 78%, and 6%, respectively at 24 h. Based on amino acid composition shifts, it can be seen that as the incubation progresses the individual amino acid composition more closely resembled BSA. In the Chesapeake Bay and lower Delaware Bay incubations, adequate material was available for individual products of BSA hydrolysis at nominal masses at 24 and 48 h to be collected (Table 2). At 24 h the nonpolar amino acids were dominant (67%), followed by the charged, polar groups (30%), and then the uncharged polar amino acids (3%). At 48 h this distribution shifted to 68%, 10%, and 22%, respectively. The nonpolar amino acid group remained relatively unchanged, while the polar charged group decreased, and the uncharged polar group increased over time. The low molecular weight fraction at 48 h was unique in the increase of uncharged, polar amino acids. This was the only occurrence of such an increase for this amino acid group (the rest were below 10%). By the end of the incubation period, however, the amino acid composition more closely resembled that of intact BSA (Fig. 6). In the lower Delaware Bay incubation (Table 2, Fig. 6), the NBSA molecular weight fraction collected at 12 h was similar to that seen in the marsh incubation. The composition was similar to BSA with charged polar amino acids (49%) as the dominant group, followed by nonpolar amino acids (48%), and uncharged, polar amino acids (3%). The b BSA fraction was also similar to the marsh incubation, with a distribution of 33%, 64%, and 3% for charged polar, nonpolar and uncharged polar, respectively. In this incubation, the 66 kDa peak corresponding to BSA was collected at 24, 36, and 48 h. The distribution of amino acids in this peak was different than expected for BSA, with nonpolar amino acids dominating (75%), followed by uncharged, polar amino acids (17%), and then charged polar amino acids (8%). Individual amino acids of the collected peaks did not show a shift over time towards intact BSA composition as seen in the other two incubations, but instead showed an overall Table 2 Major amino acid composition of proteinaceous material collected from size exclusion chromatography of natural water incubation samples at various time points Chesapeake Bay 24 h 24 h Lower Delaware Bay 24 h 48 h 24 h 24 h Marsh 24 h 36 h 12 h 12 h 12 h 12 h 24 h 45 kDa 29 kDa 17 kDa 12 kDa 165 kDa N66 kDa 13 kDa 66 kDa N100 kDa 66 kDa 44 kDa 5–52 kDa 64 kDa THR 2.18 PRO 5.59 GLY 26.85 ALA 17.35 VAL 27.71 LEU 4.59 ILE 3.39 PHE 2.49 ASP + ASN 3.02 GLU + GLN 4.17 LYS 2.66 1.35 3.68 21.20 13.84 4.37 4.92 1.72 2.94 18.98 19.29 7.71 3.66 3.14 18.88 19.53 5.51 11.33 4.26 5.22 12.64 12.11 3.73 21.63 4.21 29.37 6.24 14.56 4.70 4.35 1.87 2.09 4.85 6.14 3.04 2.93 16.19 10.16 6.34 5.84 3.25 2.94 13.32 24.02 11.97 2.92 2.67 24.16 13.52 4.88 5.17 2.18 2.12 11.91 19.73 10.75 3.39 6.90 26.71 15.23 4.13 5.03 3.25 2.28 13.67 16.13 3.29 21.63 4.21 29.37 6.24 14.56 4.70 4.35 1.87 2.09 4.85 6.14 1.15 4.83 32.22 4.80 4.59 6.51 1.21 4.50 14.67 20.65 4.87 4.11 3.59 18.25 12.63 6.43 4.82 1.83 6.05 8.93 20.37 13.00 3.44 5.50 18.73 19.48 8.48 4.60 1.70 2.21 11.34 16.30 8.22 6.17 2.08 33.53 20.69 21.46 1.97 1.98 1.37 5.37 2.47 2.89 15.17 18.06 6.79 3.84 16.52 9.69 5.31 4.98 1.18 6.50 11.96 For marsh incubations, peaks within the size ranges of LMW 5–52 kDa, 52–100 kDa, and N100–1000 kDa were combined. Chesapeake Bay and Lower Delaware Bay samples also contained individual peaks with nominal masses shown. The marsh incubation contained a combination of 12- and 24-h time points. Trp is lost during acid hydrolysis; Met and His were highly variable and have been excluded. ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx 100 Mole Percent Uncharged Polar 80 Marsh Non polar Charged Polar 60 40 20 0 BSA 100-1000 5-52 52-100 Abiotic Modified Product Molecular weight (kDa) Mole Percent 100 Chesapeake Bay 80 60 40 20 0 BSA 45 29 17 Abiotic Modified product molecular weight (kDa) 100 Mole Percent Delaware Bay 80 60 40 20 0 BSA 13 >66 48 Abiotic Modified product molecular weight (kDa) Fig. 6. The amino acid composition (mol%) of major protein hydrolysis products observed during BSA degradation in marsh and anoxic Chesapeake Bay waters. After 12 h, modified products were isolated by size exclusion chromatography, followed by amino acid analysis. Molecular weight of products that differed were calibrated using retention times of known proteins. The protein standard and a concentrate from the abiotic incubation at 92 h are included for comparison. See Fig. 3 for illustration of the range of products observed. further increase in the charged, polar amino acids, and a further decrease in nonpolar amino acids. 4. Discussion Size-exclusion chromatography and mass spectrometry of the Patuxent River incubation confirmed that modified products of protein appear rapidly in the presence of natural microbial communities. The molecular weight changes observed are consistent with pre- 9 vious reports (Nunn et al., 2003; Pantoja and Lee, 1999), where it was observed that 2–3 amino acids appear to be initially hydrolyzed from a protein during microbial degradation, and the modified product then released into the environment for further attack. In these studies it was unclear where hydrolytic attack occurred, but it has been suggested that hydrolysis is initiated from the terminal ends of the protein (Nunn et al., 2003). An examination of the primary structure of BSA (Hirayama, 1990), shows that nonpolar amino acids are situated on the terminal ends of the protein. If the termini of a protein are sites for initial hydrolysis, then we would expect nonpolar amino acids would be preferentially lost. Rather, the present results of amino acid analysis suggest the initial loss of charged polar amino acids. Although the natural conformation of BSA in solution is not firmly established, its high solubility suggests hydrophilic, charged polar amino acids are likely to be located on the exposed face with more hydrophobic amino acids enclosed. That being the case, during hydrolysis hydrophilic amino acids would be removed first, with modified products tending to have a more hydrophobic character than the original protein. Previous work by Nguyen and Harvey (2001) has shown that proteins which have survived early diagenesis have a more hydrophobic character, allowing aggregation and the potential for preservation. This is consistent with our findings of rapid aggregation of proteinaceous material in all incubations examined soon after initial hydrolysis. The present results differ from previous studies with longer experimental periods and less frequent sampling, which generally found little selective removal of amino acids during protein decay (Henrichs and Farrington, 1987; Nguyen and Harvey, 1997; Rosenfeld, 1979; Wakeham et al., 1984). The preservation of glycine and serine seen in the present results has been previously observed, but has been attributed to the refractory nature of diatom cell walls (Hecky et al., 1973; Lee and Cronin, 1984; Siezen and Mague, 1978). Carlson et al. (1985) have also suggested that binding of certain amino acids to macromolecular organic matter may also permit glycine enrichment. In the present incubations, a clear shift was apparent at the outset of incubations as polar amino acids were preferentially used, yet as the incubation progressed the overall composition shifted to more closely resemble the amino acid distribution found in the added intact protein. This suggests that while there can be selectivity in the earliest stages of hydrolysis, such selectivity is of short duration as the remaining modified proteins continue on the path of mineralization. In the case of BSA, the altered amino ARTICLE IN PRESS 10 L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx acid composition over short incubation times were transient, with longer incubation times and more degraded products, reflecting an amino acid composition similar to that of the intact protein. Given that the total amino acid composition of many proteins is similar, the analysis of particles or dissolved materials which contain both intact and modified products might be expected to show only subtle differences in amino acid composition. Many studies have examined kinetic rates of bacterial utilization of proteins and amino acids and consequently, protein degradation (Hollibaugh and Azam, 1983; Keil and Kirchman, 1991b, 1993; Samuelsson and Kirchman, 1990; Kirchman and Hodson, 1984). Even so, the number of individual proteins identified in aquatic environments is extremely limited (Tanoue, 1995; Tanoue et al., 1996; Saijo and Tanoue, 2004) and the enzymatic mechanism(s) for the utilization of protein in these environments are not well understood. The results here and many others (e.g. Harvey et al., 1995) support the rapid bacterial degradation of proteins, with BSA in the present work reduced by over half within the first 6 h and the majority removed by 12 h in all experimental replicates. Yet even over these short time scales, new proteinaceous intermediates were observed, having both greater and lower masses than the protein added. A number of these products appear to be related to BSA by sequential loss of amino acids (Fig. 4). Our results also suggest that as protein is hydrolyzed, intermediates released back into the environment are available for further degradation. These results are consistent with recent work in sediments suggesting that not all intermediate products formed from protein degradation are immediately taken up, but instead some were released back to the water (Nunn et al., 2003). Vetter et al. (1998) suggested a similar model in which proteins in sediments are partially hydrolyzed and the remaining product released. The enzymatic mechanisms of protein degradation in natural waters are important in determining rates and patterns of dissolved protein degradation. Several models exist that describe the bacterial degradation of protein, including those suggesting that extracellular enzymes are released to hydrolyze proteins (Mayer et al., 1995). The small molecular weight changes observed in the protein incubations here are consistent with extracellular hydrolysis, indicative of the action of an exoprotease, which hydrolyze amino acids from the termini of proteins (Vetter et al., 1998). The results here show that while BSA was degraded rapidly through the experimental incubation, some high molecular weight products were also formed. The pres- ence of aggregated high molecular weight products have been observed during protein degradation (Nguyen and Harvey, 2001) and hydrophobic interactions and/or hydrogen bonding have been suggested as possible mechanisms for its preservation. Additional modifications may occur under abiotic conditions (Keil and Kirchman, 1993), but are observed on longer temporal scales. Comparisons of the three estuarine sites suggest that environmental differences and bacterial community structure had little effect on the production of protein degradation products. Although there were variations among the intermediate products produced, all incubations showed a rapid decrease in added protein, indicating that differences in the bacterial community had little effect on relative degradation rates. In parallel work, Harvey et al. (in press) has observed that the addition of BSA to natural water incubations at related sites showed no consistent shifts in bacterial communities based on 16S RNA phylogenetic analysis (FISH). (Cottrell and Kirchman, 2000) have shown similar results in the examination of DOM utilization and bacterial community shifts in estuarine and coastal environments in that utilization did not correlate with shifts in the abundance of specific bacterial groups. This may not be the same for other classes of organic carbon. For example, Arnosti (2004) has discussed the case for polysaccharides; detailing experimental evidence that initial enzymatic hydrolysis is not always the slow step in subsequent degradation. Enzyme production (or its limitation) and varied extracellular enzyme distribution were also suggested as factors which might regulate organic matter degradation. These varied responses suggest that while there appears sufficient breadth among the microbial communities in these diverse environments to utilize added proteins and their products, this might not be the case for other fractions of organic matter. Although overall protein degradation was rapid, the pattern of high molecular weight products seen was not uniform among the three environments (Fig. 3). If dissolved protein complexes with organic matter creating more refractory, high molecular weight material, we would speculate that the marsh incubation would have the highest aggregation rates given its high DOM content, followed by the Chesapeake Bay and lower Delaware Bay with the lowest organic content. The marsh did show the greatest relative abundance of high molecular weight proteinaceous material; by 48 h the losses were similar to the other sites. The organic matter content of the environments may influence the degree of overall aggregation into high molecular weight pro- ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx teinaceous material, however further studies are required to investigate what effect this may have on the potential for preservation. BSA is a unique protein in that it has a somewhat different amino acid composition than many aquatic proteins (Nguyen and Harvey, 1994b), and together with commercial availability of the highly purified protein allowed small changes in molecular weight modification to be followed in the present study. The amino acid composition shifts seen in BSA were most similar in the marsh and the Chesapeake Bay incubations. Nevertheless, in all three sites modified products more closely resembled the amino acid composition of BSA as the incubation progressed, despite initial shifts in functional groups. This indicates a selective attack on the protein with the initial removal of charged polar amino acids and retention of nonpolar amino acids, specifically glycine and alanine. Although the natural conformation of BSA is not known, we can speculate that these charged polar amino acids occupy the outer shell of BSA when it is in its natural hydrated conformation. After these amino acids are removed and the interior of the protein is exposed, remaining amino acids are utilized. Support for this hypothesis can be seen in the products collected over time suggesting that charged polar amino acids are depleted and there is a corresponding increase in glycine, indicating preservation. The higher molecular weight intermediates (N 100 kDa) observed were very similar in amino acid composition to BSA. This material may contain significant amounts of native BSA or that with minor alterations caused by small changes in the chemical properties of BSA through degradation. It is interesting to note that different phytoplankton have all been found to have very similar amino acid compositions (Cowie and Hedges, 1992; Nguyen and Harvey, 1994b). Dissolved combined amino acids in various waters all had very similar amino acid compositions, including the Delaware estuary (Keil and Kirchman, 1993), particulate matter from surface oceanic and coastal waters (Siezen and Mague, 1978), Peru coastal upwelling waters (Wakeham et al., 1984), and particulate organic matter in Arctic and Antarctic waters (Hubberten et al., 1995). Although similar amino acid compositions exist in most environments due to similarity in amino acid protein composition, attack by bacteria may first require selective attack of external amino acids, namely the charged, polar amino acids. Once the interior of the proteins is exposed, all amino acids are used at equal rates. 11 5. A conceptual model of protein structure during degradation While information on specific protein modification is abundant in the biomedical literature (e.g. signaling proteins or activation) the degradation process is less specific. As a soluble, globular protein, it requires that the hydrophilic (polar, charged) amino acids of BSA are largely located on the exterior. The hydrophobic, nonpolar amino acids are directed towards the interior and the uncharged, polar amino acids throughout the structure. As the protein is utilized, smaller, modified products are formed along with aggregated higher molecular weight material. From the amino acid analysis, it is apparent that as the protein is utilized, there is a loss of charged, polar amino acids. As these amino acids are removed, the tertiary structure of the protein is compromised and the protein might unfold, exposing the interior hydrophobic amino acids. Once the protein unfolds, there is no selective attack on amino acids with utilization limited largely by the protein conformation. There does not appear to be selective attack on uncharged, polar amino acids, supporting the suggestion that the differences in the utilization of these amino acids are selective due to amino acid availability within a variable protein conformation. This suggests that one explanation for the increased aggregation could be the exposure of the interior nonpolar amino acids. Aggregates would form through hydrophobic interactions among nonpolar amino acids again surrounded by charged polar amino acids. 6. Conclusions Although proteins are degraded rapidly in natural waters, intermediates of varied molecular weight can be observed over short time periods. Amino acid analysis of a suite of these intermediate products shows that charged, polar amino acids appear preferentially lost from the model protein BSA with an increased fraction of nonpolar amino acids present in such modified products. Over time and with additional degradation, the amino acid composition returned to that more closely resembling the amino acid composition of the intact protein. Initial products also suggest that that the protein may not be hydrolyzed from the end terminus, but instead that the charged, polar amino acids present at multiple sites on the outer surface of the protein in its natural conformation are removed. As these amino acids are removed from the protein, the interior amino acids are exposed and utilized, resulting in little selective preservation. Further examination using a range of tech- ARTICLE IN PRESS 12 L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx niques (e.g. Ostrom et al., 2000) and sequencing of specific products is needed to identify amino acid locations in natural conformations of proteins to verify the initial sites of hydrolysis in the environment. Acknowledgements This work was supported by the Chemical Oceanography program of the NSF. Rachael Rearick is thanked for technical assistance with protein incubations and Angela Squier for advice and manuscript comments. Constructive comments by two anonymous reviewers substantially improved the final manuscript. Contribution No. 3926, University of Maryland Center for Environmental Science. References Arnosti, C., 2004. Speed-bumps and barricades in the carbon cycle: substrate structural effects on carbon cycling. Mar. Chem. 92, 263 – 273. Carlson, D.J., Mayer, L.M., Brann, M.L., Mague, T.H., 1985. Binding of monomeric organic compounds to macromolecular dissolved organic matter in seawater. Mar. Chem. 16, 141 – 153. Cottrell, M., Kirchman, D.L., 2000. Natural assemblages of marine proteobacteria and members of the Cytophaga–Flavobacter cluster consuming low and high molecular weight dissolved organic matter. Appl. Environ. Microbiol. 66, 1692 – 1697. Cowie, G.L., Hedges, J.I., 1992. Improved amino acid quantification in environmental samples: charged-matched recovery standards and reduced analysis time. Mar. Chem. 37, 223 – 238. Harvey, H.R., Tuttle, J.H., Bell, J.T., 1995. Kinetics of phytoplankton decay during simulated sedimentation: changes in biochemical composition and microbial activity under oxic and anoxic conditions. Geochim. Cosmochim. Acta 59, 3367 – 3377. Harvey, H.R., Dyda, R.Y., Kirchman, D.L., in press. The impact of DOM composition on bacterial lipids and community structure in estuaries. Aq. Micro. Ecol. Hecky, R.E., Mopper, K., Kilham, P., Degens, E.T., 1973. The amino acid and sugar composition of diatom cell walls. Mar. Biol. 19, 323 – 331. Hedges, J.I., 1978. The formation and clay mineral reactions of melanoidins. Geochim. Cosmochim. Acta 42, 69 – 76. Hedges, J.I., Hare, P.E., 1987. Amino acid adsorption by clay minerals in distilled water. Geochim. Cosmochim. Acta 51, 255 – 259. Henrichs, S.M., Farrington, J.W., 1987. Early diagenesis of amino acids and organic matter in two coastal marine sediments. Geochim. Cosmochim. Acta 51, 1 – 15. Hirayama, 1990. Rapid confirmation of the primary structure of bovine serum albumin by ESI/MS and FRIT-FAB LC/MS. Biochem. Biophys. Res. Comm. 173, 639. Hollibaugh, J.T., Azam, F., 1983. Microbial degradation of dissolved proteins in seawater. Limnol. Oceanogr. 28, 1104 – 1116. Hubberten, U., Lara, R.J., Kattner, G., 1995. Refractory organic compounds in polar waters: relationship between humic substances and amino acids in the Arctic and Antarctic. J. Mar. Res. 53, 137 – 149. Keil, R.G., Kirchman, D.L., 1991a. Contribution of dissolved free amino acids and ammonium to the nitrogen requirements of heterotrophic bacterioplankton. Mar. Ecol. Prog. Ser. 73, 1 – 10. Keil, R.G., Kirchman, D.L., 1991b. Dissolved combined amino acids in marine waters as determined by a vapor-phase hydrolysis method. Mar. Chem. 33, 243 – 259. Keil, R.G., Kirchman, D.L., 1993. Dissolved combined amino acids: chemical form and utilization by marine bacteria. Limnol. Oceanogr. 38 (6), 1256 – 1270. Keil, R.G., Montluçon, D.B., Prahl, F.G., Hedges, J.I., 1994. Sorptive preservation of labile organic matter in marine sediments. Nature 370, 549 – 552. Kirchman, D.L., Hodson, R., 1984. Inhibition by peptides of amino acid uptake by bacterial populations in natural waters: implications for the regulation of amino acid transport and incorporation. Appl. Environ. Microbiol. 47, 624 – 631. Kirchman, D.L., Henry, D.L., Dexter, S.C., 1989. Adsorption of proteins to surfaces in seawater. Mar. Chem. 27, 201 – 217. Knicker, H., Hatcher, P.G., 1997. Survival of protein in an organicrich sediment: possible protection by encapsulation in organic matter. Naturwissenschaften 84, 231 – 234. Lee, C., Cronin, C., 1984. Particulate amino acids in the sea: effects of primary productivity and biological composition. J. Mar. Res. 42, 1075 – 1097. Lee, S.H., Ruckenstein, E., 1988. Adsorption of proteins onto polymeric surfaces of different hydrophilicities — a case study with bovine serum albumin. J. Colloid Interface Sci. 125 (2), 365 – 379. Lourenço, S.O., Barbarino, E., Lanfer Marquez, U.M., Aidar, E., 1998. Distribution of intracellular nitrogen in marine microalgae: basis for the calculation of specific nitrogen-to-protein conversion factors. J. Phycol. 34, 798 – 811. Mannino, A., Harvey, H.R., 2000. Biochemical composition of particles and dissolved organic matter along an estuarine gradient: sources and implications for DOM reactivity. Limnol. Oceanogr. 45, 775 – 788. Mayer, L.M., Schick, L.L., Sawyer, T., Plante, C.J., 1995. Bioavailable amino acids in sediments: a biomimetic, kinetics-based approach. Limnol. Oceanogr. 40, 511 – 520. McCarthy, M., Pratum, T., Hedges, J., Benner, R., 1997. Chemical composition of dissolved organic nitrogen in the ocean. Nature 390, 150 – 154. Nagata, T., Kirchman, D.L., 1997. Roles of submicron particles and colloids in microbial food webs and biogeochemical cycles within marine environments. In: Jones, T. (Ed.), Advances in Microbial Ecology. Plenum, New York, pp. 81 – 103. Nagata, T., Fukuda, R., Koike, I., Kogure, K., Kirchman, D.L., 1998. Degradation by bacteria of membrane and soluble protein in seawater. Aquat. Microb. Ecol. 14, 29 – 37. Nguyen, R.T., Harvey, H.R., 1994a. A rapid microscale method for the extraction and analysis of protein in marine samples. Mar. Chem. 45 (1–2), 1 – 14. Nguyen, R.T., Harvey, H.R., 1994b. A rapid micro-scale method for the extraction and analysis of protein in marine samples. Mar. Chem. 45, 1 – 14. Nguyen, R.T., Harvey, H.R., 1997. Protein and amino acid cycling during phytoplankton decomposition in oxic and anoxic waters. Org. Geochem. 27 (3–4), 115 – 128. Nguyen, R.T., Harvey, H.R., 1998. Protein preservation during early diagenesis in marine waters and sediments, nitrogen-containing macromolecules in the bio- and geosphere. ACS Symposium Series, 27 (3–4), p. 88 – 112. ARTICLE IN PRESS L.C. Roth, H.R. Harvey / Marine Chemistry xx (2006) xxx–xxx Nguyen, R.T., Harvey, H.R., 2001. Protein preservation in marine systems: hydrophobic and other non-covalent associations as major stabilizing forces. Geochim. Cosmochim. Acta 65, 1467 – 1480. Nunn, B.L., Norbeck, A., Keil, R.G., 2003. Hydrolysis patterns and the production of peptide intermediates during protein degradation in marine systems. Mar. Chem. 83 (1–2), 59 – 73. Pantoja, S., Lee, C., 1999. Peptide decomposition by extracellular hydrolysis in coastal seawater and salt marsh sediment. Mar. Chem. 63, 273 – 291. Pantoja, S., Lee, C., Marecek, J.F., 1997. Hydrolysis of peptides in seawater and sediment. Mar. Chem. 57, 25 – 40. Pennock, J.R., Sharp, J.H., 1986. Phytoplankton production in the Delaware estuary: temporal and spatial variability. Mar. Ecol. Prog. Ser. 34, 143 – 155. Ostrom, P.H., Schall, M., Gandhi, H., et al., 2000. New strategies for characterizing ancient proteins using matrix-assisted laser desorption ionization mass spectrometry. Geochim. Cosmochim. Acta 64, 1043 – 1050. Rosenfeld, J.K., 1979. Amino acid diagenesis and adsorption in nearshore anoxic sediments. Limnol. Oceanogr. 24 (6), 1014 – 1021. Saijo, S., Tanoue, E., 2004. Characterization of particulate proteins in Pacific surface waters. Limnol. Oceanogr. 49, 953 – 963. Sharp, J.H., Culberson, C.H., Church, T.M., 1982. The chemistry of the Delaware estuary. General considerations. Limnol. Oceanogr. 27, 1015 – 1028. 13 Samuelsson, M.O., Kirchman, D.L., 1990. Degradation of adsorbed protein by attached bacteria in relationship to surface hydrophobicity. Appl. Environ. Microbiol. 56, 3643 – 3648. Siezen, R.J., Mague, T.H., 1978. Amino acids in suspended particulate matter from oceanic and coastal waters of the Pacific. Mar. Chem. 6, 215 – 231. Silfer, J.A., Engel, M.H., Macko, S.A., Jumeau, E.J., 1991. Stable carbon isotope analysis of amino acid enantiomers by conventional isotope ratio mass spectrometry and combined gas chromatography / isotope ratio mass spectrometry. Anal. Chem. 63, 370 – 374. Tanoue, E., 1995. Detection of dissolved protein molecules in oceanic waters. Mar. Chem. 51, 239 – 252. Tanoue, E., Ishii, M., Midorikawa, T., 1996. Discrete dissolved and particulate proteins in oceanic waters. Limnol. Oceanogr. 41, 1334 – 1343. Vetter, J.W.D., Jumars, P.A., Krieger-Brockett, B.B., 1998. A predictive model of bacterial foraging by means of freely released extracellular enzymes. Microb. Ecol. 36 (75–92). Wakeham, S.G., Lee, C., Farrington, J.W., Gagosian, R.B., 1984. Biogeochemistry of particulate organic matter in the oceans: results from sediment trap experiments. Deep-Sea Res. 31, 509 – 528.