Hydrogen isotope fractionation in algae: III. Theoretical interpretations

Hydrogen isotope fractionation in algae: III. Theoretical interpretations

Organic Geochemistry 75 (2014) 1–7 Contents lists available at ScienceDirect Organic Geochemistry journal homepage: www.elsevier.com/locate/orggeoch...

536KB Sizes 2 Downloads 81 Views

Organic Geochemistry 75 (2014) 1–7

Contents lists available at ScienceDirect

Organic Geochemistry journal homepage: www.elsevier.com/locate/orggeochem

Hydrogen isotope fractionation in algae: III. Theoretical interpretations Zhaohui Zhang a,⇑, Daniel B. Nelson b, Julian P. Sachs b a b

State Key Laboratory for Mineral Deposits Research (Nanjing University), College of Earth Sciences and Engineering, Nanjing University, Nanjing 210093, China University of Washington, School of Oceanography, Box 355351, Seattle, WA 98195, USA

a r t i c l e

i n f o

Article history: Received 7 July 2013 Received in revised form 26 May 2014 Accepted 27 May 2014 Available online 18 June 2014 Keywords: Hydrogen isotope fraction Lipid biomarkers Algae Hydrogen tunneling

a b s t r a c t Hydrogen isotope measurements of lipid biomarkers preserved in sediments are most commonly interpreted as qualitative, rather than quantitative indicators of paleoprecipitation owing to an imperfect knowledge of all factors controlling the isotopic fractionation occurring during biosynthesis. Here, we first offer a brief review of appropriate procedures for preparing enriched isotope substrates for use in tracer studies and outline the approximate dD threshold at which this transition occurs. We then present new interpretations to explain deviations from common stable isotope effects observed in our previous culture experiments and other studies. We draw particular attention to the disagreement between intercept and slope for product–substrate relationships from those predicted for isotope systems, even when R2 values are high, and attribute it to kinetic isotope fractionation. We demonstrate that reconstructing paleoenvironmental water dD values by simply adding a D to measured biomarkers dD values will result in a bias toward deuterium enriched values. This applies even to implicit reconstructions in the form of qualitative interpretations of measured lipid dD values as indicators of past hydroclimate. We therefore recommend reconstructing water dD values from lipid dD values using fractionation factor (a). We also discuss the apparently contradictory increase in D/H fractionation observed at elevated temperature and suggest that this may be the result of the unique wave-particle duality of hydrogen isotopes, which permits isotopologues to avoid surmounting the activation energy barrier that is necessary in traditional kinetic reactions. Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Improved understanding of the factors that control the hydrogen isotopic composition of algal lipid biomarkers in lacustrine and marine sediments holds great potential for direct reconstruction of water dD values, a long-standing goal of paleoclimatology. All hydrogen in phytoplankton derives from the water in their immediate environment and algae are not affected by processes that alter dD values in land plants (e.g. transpiration). Lipid biomarkers are generally well preserved in sediments and have thus become attractive targets for studies aimed at reconstructing paleoprecipitation (e.g. Sauer et al., 2001; Huang et al., 2002; Englebrecht and Sachs, 2005; van der Meer et al., 2007; Sachs et al., 2009; Smittenberg et al., 2011). As in previous papers (Zhang and Sachs, 2007; Zhang et al., 2009), the defined reference standard for hydrogen isotope ratios, described as the relative abundances of deuterium (D; 2H) and

⇑ Corresponding author. Tel.: +86 25 8359 2024; fax: +86 25 8368 6016. E-mail address: [email protected] (Z. Zhang). http://dx.doi.org/10.1016/j.orggeochem.2014.05.015 0146-6380/Ó 2014 Elsevier Ltd. All rights reserved.

protium (H; 1H) is Vienna Standard Mean Ocean Water (VSMOW). D/H is used interchangeably with R, the raw isotope ratio, and dD is defined as: dD = (Rsample/RVSMOW  1)  1000‰. The fractionation factor, a, is defined as: alipid–water = Rlipid/Rwater=(dDlipid + 1000)/ (dDwater + 1000) and the enrichment factor, e, is: elipid–water = (alipid–water  1)  1000. The hydrogen isotopic difference between a biomarker and water is defined as: Dlipid–water = dDlipid  dDwater. The large mass difference between H and D results in exceptionally large fractionation during hydrogen isotope reactions. In an extreme case, the rate constants for dissociation of H2 ? 2H and D2 ? 2D differ by 31 times, i.e. kH/kD  31 (Chang, 2005). Although for reactions where most hydrogen is bound to carbon atoms, the observed ratio of KC–H/KC–D is smaller ( 5) but still appreciable (Chang, 2005). This leads to a large range of dD values and accentuates nonlinearities of d notation that can often be neglected for other stable isotope systems. In this study we examine the disagreement between product and substrate dD values when viewed in the context of established stable isotope fractionation theory, as well as inexplicably larger fractionation at higher growth temperature. We illustrate these issues using results from two previous papers (Zhang and Sachs,

2

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7

2007; Zhang et al., 2009), in which we showed that dD values of algal lipids are recorders of water D/H ratios with high fidelity and that hydrogen isotope fractionation in algae is influenced by temperature, growth rate and species type. We also include a brief review of appropriate procedures for preparing enriched isotope substrates for use in tracer studies where both dD and R notation become inappropriate and outline the approximate dD threshold at which this transition occurs.

2. Isotope calculations in tracer-level studies As shown through decades of research using stable isotopes at artificially high concentrations in tracer studies, as d values and D/H ratios reach extremes the approximate nature of this terminology becomes inadequate to accurately predict the isotopic composition of mixtures using a mass balance approach (e.g. Brenna et al., 1997). We offer the following brief review of this subject in the context of our batch cultures as a reminder of the need to use atom fraction, or F notation in such work, and illustrate the threshold to which d notation may be applied in hydrogen isotope systems. We used Aldrich deuterium oxide (D2O) with 99.9 atom% D and mixed with distilled water at different ratios to make five waters with different dD values. The correct approach for tracer-level isotope calculation instead uses atom fractions (F): F = [D/(D + H)] = (D/H)/ [1+(D/H)] (Hayes, 1984). We compared the difference between linear mixing model calculations based on D/H ratios and F notation by preparing two water samples with each approach. They predicted dD values of 33‰ and 357‰, a difference of 324‰, but the measured difference of samples prepared based on the D/H calculation was only 1‰. The difference between the F notation samples was 313‰, and therefore within the intended range. The range over which dD notation is appropriate is illustrated in Fig. 1A. We plot the dD in log scale to define the applicable range to use dD notation in linear mixing calculations (Fig. 1A). Log (dD) vs. D/(D + H) ratio (F) begins to curve after log (dD) > 4, i.e., dD > 10,000 (Fig. 1A). The range for linear mixing calculations should thus be limited to dD < 10,000‰ where D/(D + H) = 0.1710 or D/H = 0.1713 (Fig. 1B), which more than covers the natural range of dD values.

3. Disagreement between slope and intercept on regression lines In organic geochemistry, hydrogen isotope fractionation is typically treated in the same manner as carbon and oxygen isotope fractionations. In this traditional view, the slope and intercept in linear regression equations from field observations or laboratory cultures are considered as the fractionation factor (a, slope) and the enrichment factor (e, intercept). Isotopic fractionation is described by a fractionation factor, a, where a = Rproduct/Rsubstrate. By this definition, a can be determined from any single pair of product and substrate dD measurements. From the definition dDA = (RA/Rstd  1)1000‰, it follows that dDproduct = aproduct–substrate  dDsubstrate + (aproduct–substrate  1)  1000. As a concrete example of this phenomenon, we consider the fractionation of hydrogen isotopes by photoautotrophs (algae or higher plants) during photosynthesis. Because water is the source of all hydrogen in photoautotrophs, the isotopic compositions of their lipids and environmental water are expected to be related by dDproduct = a  dDsubstrate + (a  1)  1000, in which a single fractionation factor represents the net effect of all biosynthetic processes. If multiple lipid–water pairs are measured for the same isotope reaction over a range of substrate dD values they define a regression line and the slope and intercept of this line gives two additional estimations of the fractionation factor (Table 1). The slope, alipid–water, and the intercept (a  1)  1000 (also referred to as e) should give the same estimate for the value of a using the stated equations (Sessions and Hayes, 2005), and both should equal a as calculated from any individual product–substrate pair. For well constrained isotope reactions, such as equilibrium fractionation between water and vapor, the expectation that the intercept, e, is exactly equal to (a  1)  1000, for slope-derived a, holds even for multiple consecutive isotope reactions (Craig and Gordon, 1965). Either a or e may be used to describe the isotopic effect of a reaction or sequence of reactions. Several authors have attempted to calibrate the magnitude of D/H fractionation in freshwater algae by measuring values of dD for lipids and lake water from sites covering a latitudinal range. Results are summarized in Table 2. These studies show that for biomarker hydrogen isotope data, estimates of fractionation derived from the slope compared to the estimates from the intercept of a regression line are not equal

Fig. 1. The applicable range of a binary linear mixing equation using dD notation. (A) The relation between log (dD) and D/(D + H) ratio (F). When log (dD) > 4, i.e. dD > 10,000‰ or D/(D + H) > 0.001710, the line begins to curve, i.e. any further small change of D/(D + H) ratio would lead to much larger variation in dD notation. (B) The range of applicability of dD or D/H ratio for binary linear mixing calculation. Mixing that involves an end member with very large dD values should not use dD notation for the linear mixing calculation, but rather D/(D + H) or F-notation, which is always correct.

3

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7 Table 1 The relations among biomarker and water dD values in cultures of Botryococcus braunii, Titicaca strain. Cultures

Water dD at harvest

C18:1 fatty acid dD

C29:2 alkadiene dD

DFA–W

Dalkadiene-FA

aFA–W

aalkadiene-W

A2W1 A2W2 A2W3 A2W4 A2W5

26 129 221 321 500

206 76 7 73 207

245 120 58 7 131

180 205 228 249 293

39 44 51 65 76 a average Std. dev

0.816 0.819 0.813 0.812 0.805 0.813 0.005

0.775 0.779 0.772 0.762 0.754 0.769 0.010

C18:1 fatty acid vs.water regression line: dDFA = 0.783  dDW  181. C29:2 alkadiene vs. water regression line: dDalkadiene = 0.708  dDW  219. C29:2 alkadiene vs. C18:1 fatty acid regression line: dDalkadiene = 0.905  dDFA  55.

Table 2 Calibrations between dD values of biomakers and environmental water.

a b

Analyte

Derived relationship

aslopea

ainterceptb

R2

Reference

Bulk lipids Phytoplantonic sterols Palmitic acid C18:1 fatty acid C29:2 alkadiene

dDlipid = 0.546  dDw  141 dDlipid = 0.748  dDw  199 dDlipid = 0.939  dDw  167 dDlipid = 0.783  dDw  181 dDlipid = 0.708  dDw  219

0.546 0.748 0.939 0.783 0.708

0.859 0.801 0.833 0.819 0.781

0.618 0.965 0.894 0.999 0.998

Sternberg (1988) Sauer et al. (2001) Huang et al. (2002) Zhang and Sachs (2007) Zhang and Sachs (2007)

Fractionation factor calculated from the slope of the derived relationship. Fractionation factor calculated from the intercept of the derived relationship.

D of algal lipids and water (‰, VSMOW)

(i.e. (a  1)  1000 – e) (Table 2; Sternberg, 1988; Sauer et al., 2001; Huang et al., 2002; Sessions and Hayes, 2005; Zhang and Sachs, 2007). In the most extreme example from our data, the dD of C29:2 alkadiene is related to growth water dD values by: dDalkadiene = 0.708  dDw  219 (Fig. 2). The slope, a, is 0.708, but an e value calculated from this (i.e. 0.708–1)  1000 is 292‰ rather than the measured slope of 219‰ (Table 2), with a or e values calculated from individual measurement pairs falling between these two extremes. We refer to this issue as the

Water D 1:1 line

Botryococcus braunii, Titicaca strain

500 Water

400

Reconstruction based on constant

C18:1FA C29:2alkadiene

300

y =0.783 x -181.0 (C18:1 fatty acid)

200

y =0.708 x -219.0 ( C29:2 alkadiene)

100 0 -100 -200 -300

increases with substrate D

0

100

200

300

400

500

D of waters at harvest ( ‰, VSMOW) Fig. 2. The relations among C29 alkadiene, C18:1 fatty acid, and medium water dD values in culture Botryococcus braunii, Titicaca strain, demonstrating that the absolute value of dD difference between algal lipids and water is related to the water dD value. Data are from Zhang and Sachs (2007). Triangles, solid circles, and squares represent dD values of water, C18:1 fatty acid and C29 alkadiene, respectively. The dotted arrows represent the dD difference between C18:1 fatty acid and water (DFA–W = dDfatty acid  dDW). Solid arrows represent dD differences between C29 alkadiene and C18:1 fatty acid (Dalkadiene-FA = dDalkadiene  dDfatty acid), which is the precursor for C29 alkadiene synthesis. The dashed line represents the hypothetical reconstructed water dD values by adding a ‘‘fractionation constant’’ (D, the smallest one among five cultures) to the measured biomarker dD values. Such a reconstruction would dampen the signal of relatively deuterium-enriched water in the paleoenvironment.

‘‘slope/intercept disagreement’’. The fractionations should be equivalent, and the fact that they are not implies that the relationship is not as simple as has been assumed. Sessions and Hayes (2005) suggested that slope/intercept disagreement might largely be attributable to the relatively low R2 values in the data sets that they analyzed, but our data demonstrate the slope/intercept disagreement despite R2 values that are in excess of 0.99 (Fig. 2). This suggests that the issue is a real and unexplained phenomenon and requires explanations. Lipid biosynthesis reactions are complex and typically involve multiple equilibrium and kinetic isotope effects (Hayes, 2001). Because it is usually not possible to measure all precursor compounds and intermediate products, intermediate fractionation steps are commonly grouped into one ‘apparent fractionation factor’ that is defined as a single a or e value. Use of this convention to describe the isotope effects in lipid biosynthesis might therefore be a possible cause of the slope/intercept disagreement. If all hydrogen in photoautotrophic lipids is fundamentally derived from water, one view of biosynthesis is to treat it as a black box with H in water as input and H in the biomarkers as output. For algae, if such fractionation is in thermodynamic equilibrium, the pairs of biomarkers and water would fall on a line: dDlipid = alipid–water  dDwater + (alipid–water  1)  1000. However, taking this view results in the previously discussed slope/intercept disagreement. As an alternative approach to address the hydrogen isotope ratios of lipids during biosynthesis, consideration in the context of a kinetic isotope model (e.g. Criss, 1999) may more adequately capture the complexity and allow room for a plausible explanation. Generalized isotope fractionation during algal growth can be written as ak

AD þ WH AH þ WD k

where A and W stand for a particular algal biomarker and water, respectively, H and D for hydrogen and deuterium, respectively. The biosynthesis involves multiple steps from water and here we treat the whole process as one box with input and output shown only. The forward and the reverse reactions don’t proceed at identical rates, but rather at rates indicated by the quantities k and ak indicated by the arrows (k is a first-order rate constant), multiplied

4

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7

by the concentration terms. Given the fact that the amount of hydrogen atoms in the water was approximately 1000 times greater than that in the algal biomass, Rw can be considered a constant. Adapting from Criss (1999), the dD value of an algal lipid is: dDA ¼ ekt  dDiA þ 1000  ðaAW  1Þ  ð1  ekt Þ þ aAW  ð1  ekt Þ  dDW

where t is time, and dDiA is the initial lipid dD value at t = 0, i.e. when the inoculant was brought into culturing. This equation contains three terms. The first two are constants for five batch cultures for a given strain, but the third term is a function of the water dD values (i.e. the heavier the water dD value, the larger DA–W). Since the five cultures for each species in our experiments (Zhang and Sachs, 2007) started at the same time and were grown under the same growing conditions, ekt can be treated as a constant. In this case, the slope of the linear regression is aA–W  (1  ekt), and the intercept = ekt  dDiA + 1000  (aA–W  1)  (1  ekt). The terms aA–W and 1000  (aA–W  1), which describe the slope and intercept under thermodynamic equilibrium are still included in this equation, but additional terms are included to describe the kinetic effects. If, after infinite time equilibrium is attained, i.e. t ? 1, then, ekt ? 0 and as a result, the equation can be simplified to dDA = 1000  (aA–W  1) + aA–W  dDW, equivalent to the equilibrium condition, and under such circumstances a and e will be consistent with the observed relation between dDA and dDW. Although our experiments were conducted with D-enriched material and therefore include a range of dD values that are outside the range of available international reference materials (Sternberg, 1988; Sauer et al., 2001; Huang et al., 2002; Sessions and Hayes, 2005), we consider the fact that previous investigations also observed the slope/intercept disagreement to be an argument against the idea that our results are a referencing artifact. Similarly it could be argued that our results might be driven by slight differences in the growth curves of the algae in our batch cultures, but this would require a systematic effect as a function of the substrate dD values, which we also consider unlikely. Instead, we suggest that the disagreements in spite of high R2 values in our studies and previous observations in other studies most likely stem from the fact that although hydrogen isotope fractionation in sedimentary organic geochemistry is most often considered in the context of thermodynamic equilibrium (e.g. Huang et al., 2002; Sessions and Hayes, 2005), in reality such equilibrium is likely never attained during biosynthetic reactions and they should be considered kinetic. 4. Possible bias in paleoprecipitation reconstruction when using D Although most hydrogen isotope applications do not make direct use of D notation (dDproduct  dDsubstrate), paleoclimate studies that interpret lipid dD values as a direct indicator of hydrology without consideration of the magnitude of D/H fractionation are effectively doing so implicitly (Polissar and D’Andrea, 2014). Fig. 2 demonstrates how D values between lipid biomarkers and water are related with water dD values in cultures of B. braunii A race, Titicaca strain. dD values of both C18:1 fatty acids and alkadienes in five cultures were linearly correlated with water dD values at harvest (Table 1), with R2 = 0.998 and 0.999, respectively (Fig. 2). DFA–W values of 180‰, 205‰, 228‰, 249‰ and 293‰ in five cultures corresponded to water dD values of 26‰, 129‰, 221‰, 321‰ and 500‰, respectively (Table 1; Fig. 2).The magnitude of DFA–W therefore changes as water dD values change (DFA–W represented by dotted arrows); the larger dD, the bigger magnitude of DFA–W. The same effect occurs for the C29 alkadiene, but the slope of the regression line for C29 alkadiene dD–water dD (0.708) is smaller than that for C18:1 fatty acid (0.783) (Fig. 2). C29

alkadiene is the product of C18:1 fatty acid decarboxylation (Templier et al., 1984, 1991), so the progressive decrease in slope with each subsequent measured biosynthetic product reflects the D change at each stage and not just for the earliest measured product. Presumably, if another lipid were synthesized from C29 alkadiene, its dD value would show analogous substrate-dD dependence on the C29 alkadiene, resulting in an even smaller slope. This illustrates how different compounds would record different dD value changes in response to identical changes in the dD value of source water and therefore why it is important to consider the fractionation factor in interpreting hydrogen isotope data. In addition, this might also serve as a good example how multiple fractionation processes involved in a ‘‘net’’ fractionation result in stronger slope/ intercept disagreement. We therefore recommend reconstructing water dD values from lipid dD values using a, which accounts for the D/H discrimination between lipid and water, using the equation: dDw = (dDlipid + 1000)/a  1000 (Zhang and Sachs, 2007). The fractionation factor, a, can obtained through culture experiments or core-top studies. In our case, despite the increased DFA–W values with increased water dD values, the a values calculated from individual product–substrate pairs in these five cultures are nearly constant, 0.816, 0.819, 0.813, 0.812 and 0.805 (average = 0.813, standard deviation 0.005) (Table 1). Similarly, the a values of alkadienes calculated this way in these five cultures are 0.775, 0.779, 0.772, 0.762 and 0.754 (average = 0.769, standard deviation = 0.01) (Table 1). This suggests that the fractionation calculated from any one of these pairs would give a reasonable approximation of a, and therefore provide a more reliable means of comparing the change in source water dD values that is driving each lipid dD value. Interestingly, the a values calculated from product–substrate pairs agree much more closely with the intercept-derived fractionation factor from the regression equations than with those calculated from the slope. Converting each of the average a values above to e values for comparison, 187‰ and 231‰ for C18:1 fatty acid and C29 alkadiene agree much more closely with the intercepts of 181‰ and 219‰, respectively, than they do with the slope-derived e values of 217‰ and 292‰, also respectively (Table 1). This again highlights the disagreement between the slope and intercept.

5. Kinetic hydrogen isotope effect in enzyme-catalyzed reactions: rate independence In the energy profile of an uncatalyzed reaction (Fig. 3A), y is the free energy difference for the reaction (y = RT  ln Keq) and x is the activation energy for the forward direction, with x + y representing the activation energy for the reverse (Cleland and Northrop, 1999). Lipid synthesis reactions are enzyme catalyzed and therefore involve the formation of an enzymatic substrate complex, ES, between the substrate and the enzyme, and it is assumed that reactants are always in equilibrium with ES. Following formation, the ES complex isomerizes to an enzymatic product, followed by dissociation of the product to give a free enzyme again (Fig. 3B). In such a profile for an enzyme catalyzed reaction with the same Keq (Fig. 3B), the activation energy is reduced to that required for the conversion of ES to EP (z). A kinetic isotope effect (KIE) results from activation energy differences for the different isotopologue reactants, and much of its magnitude is due to the differences of zero-point energy (ZPE) between the ground state and the transition state of the reaction. KIE is the ratio of rates of the two isotopologue reactants (molecules that only differ in their isotopic compositions), e.g., KIE = kH/kD. In biochemistry, a0 is defined as a0 = Rreact/Rproduct = KIE = kH/kD (Valentine et al., 2004). In this paper, we define a as Rlipid/Rwater so that a = 1/a0 = 1/(kH/kD). For proton transfer

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7

5

Fig. 3. Reaction-coordinate diagram for the conversion of reactant (substrate, S) to product (P). (A) In a profile of an un-catalyzed reaction, if we assume y is the free energy difference for the reaction (y = RT  ln Keq) and x is the activation energy for the forward reaction, then x + y is the activation energy in the back reaction. Modified from Cleland and Northrop (1999). (B) In an enzyme-catalyzed reaction, transition state theory assumes that the enzyme and the substrate combine to form an enzymatic substrate (ES) complex, which isomerizes to enzymatic product (EP), followed by dissociation of the product (P) to give the free enzyme again. The value of y (RT  ln Keq) is the same but the activation energy is now much lowered, which is now the activation energy for the conversion of ES to EP (z), less than x in noncatalyzed reaction. Modified from Melander and Saunder (1980). (C) An example of ground-state hydrogen tunneling. The reactant well (S) is on the left and the product well (P) is on the right. The fine lines represent the probability (nuclear k2) of finding the nuclei in the reactant or the product wells. More overlap between the probability functions of the S and P results in higher tunneling probability. The superposition of wave functions for each vibrational level illustrates the increasing barrier penetration with increasing temperature. Modified from Knapp and Klinman (2002).

reactions, one of the most fundamental and prevalent processes in biology, this ratio of rates between H and D isotope is characteristic of the reaction coordinate and the nature of the transition state (Kohen, 2006):

lnðkH =kD Þ  ðDGzH  DGzD Þ=RT; i:e:kH =kD  exp½ðDGzH  DGzD Þ=RT where DGàD  DGàH  ZPEàD  ZPERD  (ZPEàH  ZPERH), where T is temperature, and R is the gas constant (Melander and Saunder, 1980; Kiefer and Hynes, 2006; Kohen, 2006). Kinetic fractionation is therefore only dependent on the differences in activation energy for the different isotopologue reactants and temperature, and is independent of substrate concentration. Accordingly, we observed that hydrogen isotope fractionation in fatty acids changed as a function of temperature in cultures of Eudorina unicocca and Volvox aureus, but not as a function of growth rate differences induced by nutrient limitation in Thalassiosira pseudonana (Zhang et al., 2009). 6. Hydrogen tunneling, a possible mechanism underlining inexplicable hydrogen isotope fractionation phenomena The algae species E. unicocca and V. aureus grown at 25 °C exhibited substantially larger D/H fractionation than those grown at 15 °C (Zhang et al., 2009), which is against the classic view that mass-dependent isotope fractionation decreases with increased temperatures. For example, the carbon isotope fractionation associated with malate synthesis in CAM plant Kalanchoe daigremontiana is temperature dependent. The fractionation is approximately 4‰ at low temperatures and approaches zero at higher temperatures (Deleens et al., 1985). Within the range of algal physiological temperatures in which our temperature experiments were conducted, isotope fractionation theory would predict a decrease in D/H fractionation with increasing temperature. Our observation of a temperature effect of opposite sign therefore requires an alternative explanation.

processes, depending on their thermal energy in relation to the obstructing barrier (Smedarchina et al., 2006). Transition-state theory (TST) considers only the particle-like properties of matter. Quantum tunneling is the penetration of a particle into a region that is excluded in classical mechanics (due to it having insufficient energy to overcome the potential-energy barrier). In some cases tunneling is thought to be the dominant process by which hydrogen transfer is accomplished (Knapp and Klinman, 2002). The wave nature of particles allows for the possibility that a particle penetrates a thin barrier even if the particle energy is less than the height of the barrier. From a classical mechanics point of view, tunneling cannot easily be explained since it would be the equivalent of a ball going through a wall without damaging the wall. A measure of this uncertainty in the position of a given hydrogen atom is the de Broglie wavelength, k = h/(2mE)1/2, in which h is Planck’s constant, m the mass of the particle, and E its energy. k2 is the probability that a particle will be in a given region of space (Fig. 3C). Assuming E = 20 kJ/mol ( 5 kcal/mol) the de Brogilie wavelengths are 0.63 Å and 0.45 Å for protium (H) and deuterium (D), respectively (Knapp and Klinman, 2002). As hydrogen is typically transferred over a similar distance (< 1 Å), that is close enough to the hydrogen wavelength for tunneling to occur. In this dynamic view of enzyme catalysis it is the width, and not the height (as with TST), of an energy barrier that controls the reaction rate (Sutcliffe and Scrutton, 2000; Fig. 3C). The separation distance and symmetry of the S (substrate) and P (product) wells control tunneling probability (Fig. 3C). All else being equal, hydrogen has a higher tunneling probability than deuterium since a heavy isotope has a lower ZPE and its probability function is more localized in its well (Kohen, 2006). Additionally, all regions of the enzyme (not just the active site) likely contribute to the vibrations that drive quantum tunneling, thus providing a possible reason why enzymes are much larger than the active site alone (Sutcliffe and Scrutton, 2002). Quantum tunneling is thus an attractive means of transferring hydrogen from reactant to product in enzyme-catalyzed reactions where classical over-the-barrier energy requirements are large.

6.1. Review on hydrogen tunneling Hydrogen isotopes are distinguished from other stable isotopes due to the wave-particle duality of matter. Hydrogen atoms and protons can transfer not only by classical over-barrier processes as described above, but also by quantum mechanical through-barrier

6.2. Temperature dependence of enzyme vibration in hydrogen tunneling and the effect on biomarker isotope fractionation Increased hydrogen isotope fractionation at higher growth temperature is not explained by classic isotope theory. Basran et al.

6

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7

(2001) observed a similar temperature-dependence and concluded that hydrogen tunneling may be significant at physiological temperatures, or even the sole means by which enzymes catalyze hydrogen transfer during C–H bond breakage. Kohen et al. (1999) measured thermophilic alcohol dehydrogenase (ADH) in the bacterium Bacillus stearothermophilus over a very wide temperature range (5–65 °C) and showed that tunneling made a significant contribution at 65 °C where the thermophilic enzyme was optimally functional. Decreased tunneling was observed below room temperature, indicating a transition for the protein at approximately 30 °C, analogous to previous findings with mesophilic ADH at 25 °C (Cha et al., 1989). This decrease in tunneling at reduced temperature was ascribed to reduced protein mobility (Kohen et al., 1999). Kohen (2006) put forth a model in which an enzyme’s vibrations facilitated tunneling by reducing the distance over which hydrogen must travel between source and target molecules. In theoretical calculations, tunneling distance in a reaction catalyzed by the soybean enzyme lipoxygenase seems to be shorter than the distance between the source and target molecules (Hatcher et al., 2004). This suggests that whole-enzyme vibrations are needed to bring the two molecules closer when tunneling occurs. At reduced temperatures the amplitude of promoting vibrations may decrease below a level required for effective tunneling and the reaction begins to approximate more classical behavior with minimal tunneling (Kohen et al., 1999), i.e. fractionation observed will be mostly from normal KIE. We suggest that hydrogen tunneling effects may account for the observed increased D/H fractionation at higher temperatures in lipids from E. unicocca and V. aureus (Zhang et al., 2009). Hydrogen tunneling effects may also explain similar seemingly anomalous temperature effects in methane oxidizing bacteria, where a more than a two fold increase in D/H fractionation was observed for cultures grown at 26 °C as compared to those grown at 11 °C, opposite the expected decrease in the magnitude of kinetic isotope effects as temperature increases (Coleman et al., 1981).

in our dDlipid  dDwater regressions and previous studies is not explained by thermodynamic isotope fractionation often incurred in organic geochemistry studies. We attribute it to kinetic isotope fractionation. We reviewed how D (dDbiomarker  dDwater) is related to water dD values, and that this difference is especially apparent in hydrogen isotope studies due to the large isotopic fractionation that occurs. Therefore, reconstructing paleoenvironmental water dD values by simply adding a D to measured dD values of biomarkers will result in a bias toward deuterium-enriched values. We therefore recommend reconstructing water dD values from lipid dD values using a, which could be obtained through culture experiments or core-top studies. We also show that the kinetic isotope effect in an enzyme catalyzed reaction is dependent on temperature and the biosynthesis pathway, but is seemingly rate independent. The substantially larger hydrogen isotope fractionation at higher growth temperature is not explained by traditional isotope theory. We offer an alternative hypothesis of hydrogen tunneling, which may contribute to the temperature dependence of D/H fractionation in algal lipid synthesis, with increased hydrogen tunneling at elevated temperature resulting in increased fractionation.

7. Discussion and future studies

References

We discussed possibilities responsible for the disagreement between slope and intercept in our culture experiments and previous studies. We suggest that kinetic fractionation can offer some explanation because biosynthesis most likely does not reach thermodynamic equilibrium. One way to test the kinetic hypothesis is to set up a series of culture experiments with different water dD values but grown under the same conditions, and to frequently sample the algae and measure its biomarker dD values. As kinetic fractionation evolves with time, we shall see the change of algal biomarker dD values with time. Correspondingly the disagreement between slope and intercept shall shrink. Though hydrogen tunneling could explain the substantially larger fractionation at higher growth temperature, this hypothesis requires more investigation. A set of experiments involving different temperature gradients, a wider range (within the physiological temperature) and more biomarkers would be essential. Thus far there have been very few studies on the effects of temperature effect on hydrogen isotope fractionation in algal biosynthesis. Nevertheless this hypothesis could stimulate organic geochemists to design experiments and better understand the mechanism under neglected or inexplicable hydrogen isotope phenomena.

Basran, J., Sutcliffe, M.J., Scrutton, N.S., 2001. Deuterium isotope effects during carbon–hydrogen bond cleavage by trimethylamine dehydrogenase. The Journal of Biological Chemistry 276, 24581–24587. Brenna, J.T., Corso, T.N., Tobias, H.J., Caimi, R.J., 1997. High-precision continuousflow isotope ratio mass spectrometry. Mass Spectrometry Reviews 16, 227–258. Cha, Y., Murray, C.J., Klinman, J.P., 1989. Hydrogen tunneling in enzyme reactions. Science 243, 1325–1330. Chang, R., 2005. Physical Chemistry for the Biosciences. University Science Books, Sausalito, pp. 311–400. Cleland, W.W., Northrop, D.B., 1999. Energetics of substrate binding, catalysis, and product release. In: Schramm, V.L., Purich, D.L. (Eds.), Enzyme Kinetics and Mechanism. Academic Press, San Diego, pp. 3–27. Coleman, D.D., Risatti, J.B., Schoell, M., 1981. Fractionation of carbon and hydrogen isotopes by methane-oxidizing bacteria. Geochimica et Cosmochimica Acta 45, 1033–1037. Craig, H., Gordon, L.I., 1965. Deuterium and 18O variations in the ocean and the marine atmosphere. In: Tongiorgi, E. (Ed.), Proceedings of a conference on stable isotopes in oceanographic studies and paleotemperatures, Spoleto, pp. 9–130. Criss, R.E., 1999. Principles of Stable Isotope Distribution. Oxford University Press, New York. Deleens, E., Treichel, I., Otleary, M.H., 1985. Temperature dependence of carbon isotope fractionation in CAM plants. Plant Physiology 79, 202–206. Englebrecht, A.C., Sachs, J.P., 2005. Determination of sediment provenance at drift sites using hydrogen isotopes and unsaturation ratios in alkenones. Geochimica et Cosmochimica Acta 69, 4253–4265. Hatcher, E., Soudackov, A.V., Hammes-Schiffer, S., 2004. Proton-coupled electron transfer in soybean lipoxygenase. Journal of the American Chemical Society 126, 5763–5775. Hayes, J.M., 1984. Practice and principles of isotopic measurements in organic geochemistry. In: Meinschein, W. (Ed.), Organic Geochemistry of Contemporaneous and Ancient Sediments. Society of Economic Paleontologists, Bloomington, pp. 1–31. Hayes, J.M., 2001. Fractionation of the isotopes of carbon and hydrogen in biosynthetic processes. In: Walley, J.W., Cole, D.R. (Eds.), Stable Isotope Geochemistry, Reviews in Mineralogy and Geochemistry, vol. 43. Mineralogical Society of America, Washington, DC, pp. 225–278.

8. Conclusions We have attempted to address unexplained aspects of hydrogen isotope systematics in algal culture experiments. The slope (a)–intercept (e) disagreement in spite of high R2 that we observe

Acknowledgements This work was supported by the Chinese National Science Foundation under Grants NSFC 41273086 and 41173080, and by the Scientific Research Foundation for the Returned Overseas Chinese Scholars, State Education Ministry (Z.Z.). This material is based upon work supported by the U.S. National Science Foundation under Grant Numbers 1027079, and by a grant from the Gary Comer Science and Education Foundation (J.S.). We are very much grateful to Drs. Lorenz Schwark, Yoshito Chikarashi and two anonymous reviewers for their constructive reviews. Associate Editor—Lorenz Schwark

Z. Zhang et al. / Organic Geochemistry 75 (2014) 1–7 Huang, Y., Shuman, B., Wang, Y., Webb, T., 2002. Hydrogen isotope ratios of palmitic acid in lacustrine sediments record late quaternary climate variations. Geology 30, 1103–1106. Kiefer, P.M., Hynes, J.T., 2006. Interpretation of primary kinetic isotope effects for adiabatic and nonadiabatic proton-transfer reactions in a polar environment. In: Kohen, A., Limbach, H.-H. (Eds.), Isotope Effects in Chemistry and Biology. Taylor & Francis, Boca Raton, pp. 549–578. Knapp, M.J., Klinman, J.P., 2002. Environmentally coupled hydrogen tunneling. Linking catalysis to dynamics. European Journal of Biochemistry 269, 3113– 3121. Kohen, A., 2006. Kinetic isotope effects as probes for hydrogen tunneling in enzyme catalysis. In: Kohen, A., Limbach, H.-H. (Eds.), Isotope Effects in Chemistry and Biology. Taylor & Francis, Boca Raton, pp. 743–764. Kohen, A., Cannio, R., Bartoulucci, S., Klinman, J.P., 1999. Enzyme dynamics and hydrogen tunneling in a theromphilic alcohol dehydrogenase. Nature 399, 496– 499. Melander, L., Saunder, W.H., 1980. Reaction Rates of Isotopic Molecules. John Wiley & Sons. Polissar, P.J., D’Andrea, W.J., 2014. Uncertainty in paleohydrologic reconstructions from molecular dD values. Geochimica et Cosmochimica Acta 129, 146–156. Sachs, J.P., Sachse, D., Smittenberg, R., Zhang, Z.H., Battisti, D.S., Golubic, S., 2009. Southward movement of the Pacific intertropical convergence zone AD 1400– 1850. Nature Geosciences 2, 519–525. Sauer, P.E., Eglinton, T.I., Hayes, J.M., Schimmelmann, A., Sessions, A., 2001. Compound-specific D/H ratios of lipid biomarkers from sediment as a proxy for environmental and climatic conditions. Geochimica et Cosmochimica Acta 65, 213–222. Sessions, A.L., Hayes, J.M., 2005. Calculation of hydrogen isotopic fractionations in biogeochemical systems. Geochimica et Cosmochimica Acta 69, 593–597. Smedarchina, Z., Siebrand, W., Fernández-Ramos, A., 2006. Kinetic isotope effects in multiple proton transfer. In: Kohen, A., Limbach, H.-H. (Eds.), Isotope Effects in Chemistry and Biology. Taylor & Francis, Boca Raton, pp. 521–547.

7

Smittenberg, R.H., Saenger, C., Dawson, M.N., Sachs, J.P., 2011. Compound-specific D/H ratios of the marine lakes of Palau as proxies for West Pacific Warm Pool hydrologic variability. Quaternary Science Reviews 30, 921–933. Sternberg, L.D.L., 1988. D/H ratios of environmental water recorded by D/H ratios of plant lipids. Nature 333, 59–61. Sutcliffe, M.J., Scrutton, N.S., 2000. Enzymology takes a quantum leap forward. Philosophical Transactions of the Royal Society of London 358, 367–386. Sutcliffe, M.J., Scrutton, N.S., 2002. A new conceptual framework for enzyme catalysis. Hydrogen tunneling coupled to enzyme dynamics in flavoprotein and quinoprotein enzymes. European Journal of Biochemistry 269, 3096–3102. Templier, J., Largeau, C., Casadevail, E., 1984. Mechanism of non-isoprenoid hydrocarbon biosynthesis in Botryococcus braunii. Phytochemistry 23, 1017–1028. Templier, J., Largeau, C., Casadevail, E., 1991. Non-specific elongationdecarboxylation in biosynthesis of cis and transalkadienes by Botryococcus braunii. Phytochemistry 30, 175–183. Valentine, D.L., Sessions, A.L., Tyler, S.C., Chidthaisong, A., 2004. Hydrogen isotope fractionation during H2/CO2 acetogenesis: hydrogen utilization efficiency and the origin of lipid-bound hydrogen. Geobiology 2, 179–188. van der Meer, M.T.J., Baas, M., Rijpstra, W.I.C., Marino, G., Rohling, E.J., Sinnighe Damsté, J.S., Schouten, S., 2007. Hydrogen isotopic compositions of long-chain alkenones record freshwater flooding of the Eastern Mediterranean at the onset of sapropel deposition. Earth and Planetary Science Letters 262, 594–600. Zhang, Z.H., Sachs, J.P., 2007. Hydrogen isotope fractionation in freshwater algae: I. Variations among lipids and species. Organic Geochemistry 38, 582–608. Zhang, Z.H., Sachs, J.P., Marchetti, A., 2009. Hydrogen isotope fractionation in freshwater and marine algae: II. Temperature and nitrogen-limited-growth-rate effects. Organic Geochemistry 40, 428–439.