Breakdown of chlorophyll: A nonenzymatic reaction accounts for the formation of the colorless “nonfluorescent” chlorophyll catabolites

May 30, 2003
100 (12) 6910-6915

Abstract

Senescent higher plants degrade their chlorophylls (Chls) to polar colorless tetrapyrrolic Chl catabolites, which accumulate in the vacuoles. In extracts from degreened leaves of the tree Cercidiphyllum japonicum an unpolar catabolite of this type was discovered. This tetrapyrrole was named Cj-NCC-2 and was found to be identical with the product of a stereoselective nonenzymatic isomerization of a “fluorescent” Chl catabolite. This (bio-mimetic) formation of the “nonfluorescent” catabolite Cj-NCC-2 took place readily at ambient temperature and at pH 4.9 in aqueous solution. The indicated nonenzymatic process is able to account for a crucial step during Chl breakdown in senescent higher plants. Once delivered to the acidic vacuoles, the fluorescent Chl catabolites are due to undergo a rapid, stereoselective isomerization to the ubiquitous nonfluorescent catabolites. The degradation of the Chl macrocycle is thus indicated to rely on just two known enzymes, one of which is senescence specific and cuts open the chlorin macroring. The two enzymes supply the fluorescent Chl catabolites, which are “programmed” to isomerize further rapidly in an acidic medium, as shown here. Indeed, only small amounts of the latter are temporarily observable during senescence in higher plants.
Approximately 15 years ago chlorophyll (Chl) catabolism was still considered an enigma (1), in marked contrast to the detailed knowledge available on the biosynthesis of the Chls (2). As Chl metabolism is probably the most visible manifestation of life on Earth, and total annual turnover of this complex pigment is estimated to involve <1,000 million tons, the unresolved problem of Chl breakdown was rather dissatisfying (1, 3).
Over the last years, several of the mysteries of Chl catabolism have been revealed. In senescent higher plants Chl breakdown proceeds via pheophorbide a, which is rapidly degraded to colorless (nonfluorescent) tetrapyrrolic products (see Fig. 1; refs. 47). The constitution of such a colorless Chl catabolite from the monocot barley allowed the first structural insight into Chl degradation (4). Contrary to all earlier expectations, only trace amounts of (red) colored intermediate breakdown products are detectable (6). These are the products of an oxidative opening of the macrocycle of pheophorbide a (8, 9). This key step is catalyzed by a monooxygenase, whose activity is expressed specifically during senescence (10). The product of the oxidative ring cleavage step is rapidly transformed by an enzymatic reduction (11, 12) that leads to temporarily observable and faintly yellow-colored, primary fluorescent Chl catabolites (FCCs). The FCCs have been suggested to be the catabolic precursors of the colorless, nonfluorescent Chl catabolites (NCCs) (13). The NCCs appear to represent the final products of endogenous Chl breakdown in higher plants (5, 6). Surprisingly, they are found in the vacuoles of the degreened plant material rather than in chloroplasts from which they must originate (14).
Fig. 1.
Outline of the breakdown of the Chl macrocycle in higher plants. Pheophorbide a is oxidized to an enzyme-bound form of the RCC, which is reduced to FCC and then isomerized to NCC.
Despite the progress made on Chl breakdown in senescent higher plants, the catabolic means for the transformation of the temporarily observable FCCs into NCCs are still unknown (6, 8, 15). We have addressed this specific problem and report here (i) the discovery of the unpolar NCC Cj-NCC-2 (1) in extracts from degreened leaves of the tree Cercidiphyllum japonicum, (ii) the rapid, stereoselective, and nonenzymatic isomerization of the FCC Ca-FCC-2 (2) (16) to an NCC, which was identified with Cj-NCC-2 (1), and (iii) the stereochemical assignment of two of the chiral centers in natural NCCs. The isomerization of the FCC 2 to the NCC 1 was found to take place rapidly in weakly acidic solution, i.e., under conditions similar to those found in the vacuoles of senescent leaves (14, 17).

Experimental Procedures

Materials. Reagents used were reagent-grade commercial, and solvents were distilled before use. Reagents and HPLC solvents were from Fluka and Merck. Sep-Pak-C18 cartridges were from Waters.
Chromatography. Flash chromatography was done on silica gel 60 from Fluka. TLC was done on plates (0.25 mm) with silica gel 60 F254 from Merck.
HPLC. For analytical work, we used a Hypersil (Cheshire, U.K.) ODS (5 μm, 250 × 4.6 mm i.d.) as a column and a Gynkotek (Germering, Germany) M480G pump with vacuum on-line degasser. For preparative work, we used a Hypersil ODS (5 μm, 250 × 21.2 mm i.d.) and a Gynkotek M300 pump. Solvents were degassed by sonication. Detectors used were a Gynkotek diode array detector UVD 340 and a Hitachi SPD-6AV UV/visible detector. All analytical chromatograms were taken at 5°C, all preparative chromatograms were taken at room temperature, and data were processed by a Gynkotek HPLC data system and GYNKOSOFT 5.50.
Spectroscopy. To obtain UV/visible spectra we used a Hitachi-U3000 spectrophotometer at λmax(nm)/(log ε). Concentrations were calculated following Lambert-Beer's law. To obtain CD spectra we used a Jasco-J715 spectra-polarimeter at λmax(nm) and λmin(nm)/(Δε). To obtain NMR spectra we used a Varian Unityplus 500 [δ(C1HD2OD) = 3.31 ppm and δ(13CD3OD) = 49.0 ppm]. For MS we used a Finnigan MAT 95-S in (+)-ion mode. For fast atom bombardment–MS we used a cesium gun at 20 keV and a glycerol matrix. Electrospray ionization–MS was done at 3.2 kV with water/MeOH mixtures as solvents.
Senescent Leaves from C. japonicum Sieb and Zucc. Yellow leaves from C. japonicum trees were collected in late September in the botanical institutes of the University of Zürich and the University of Innsbruck and were extracted as described by Curty and Engel (18).
Preparation of Ca-FCC-2 (2). A sample of ≈70 nmol of Ca-FCC-2 (2) in 10-ml solution (MeOH/water, 30:70, vol/vol) was prepared with extracts from ripe sweet pepper (Capsicum annuum) and isolated as described by Mühlecker et al. (16). The FCC 2 was stored frozen (at —80°C) for ≈20 days before being used for the experiments described here.
Isolation of Cj-NCC-2 (1). A crude solid extract (5.8 g) was obtained essentially as described (18) from 5.25 kg (fresh weight) of yellow (senescent) leaves of C. japonicum. A total of 4.8 g of this solid extract, which contained impure Cj-NCC-1 (3) (18) and Cj-NCC-2 (1) in a ratio of ≈23:1, was subjected to flash chromatography. A sample of ≈1 g of a yellow powder was obtained, which contained still impure Cj-NCC-1 (3) and Cj-NCC-2 (1) in a ratio of ≈7:1. Flash chromatography (silica gel; eluent/dichloromethane/MeOH/water 800:200:1.5, vol/vol) of this powder yielded a fraction containing 26.1 mg of a crude sample of Cj-NCC-2 (1), free of 3 (and other fractions containing both 1 and 3 in varying amounts). A homogeneous fraction containing 1 was collected (16.5 min) from 3.6 mg of the crude 1, subjected to preparative HPLC [10 ml/min (200 mM) ammonium acetate in MeOH/water 65:35 vol/vol (pH 6.8)], and desalted with a Sep-Pak cartridge. After the solvent was evaporated in vacuo at <0°C, 0.88 mg (1.4 μmol) of pure Cj-NCC-2 (1) was obtained (see Supporting Text, which is published as supporting information on the PNAS web site, www.pnas.org, for details).
Characterization of Cj-NCC-2 (1). The sample of 0.86 mg of authentic 1 was characterized as follows: MS (nano-electrospray ionization) m/z (%), 631.2 (18), 630.2 (57), 629.2 (100, [M+H]+), 597.2 (26, [M+H-32]+), 506.2 (12, [M+H-123]+); high-resolution MS, experimental 629.2990, calculated 629.2975. UV/visible, (MeOH) λmax (log ε) 239 (sh, 4.34), 312 (4.25); CD (MeOH) λmin/max (Δε) 225 (24.0), 251 (—8.1), 258 (—7.5), 283 (—17.9), 317 (6.5) (see Fig. 2); 1H NMR (500 MHz, ≈2.3 mM in CD3OD, 26°C, tetramethylsilane) δ = 0.99 (H3C-82), 1.93 (H3C-181), 1.94 (H3C-21), 2.08 (H3C-121), 2.24 (H3C-71), 2.30–2.36 (H2C-172), 2.42 (H2C-81), 2.50 (HAC-20), 2.60–2.75 (H2C-171), 2.84 (HBC-20), 3.75 (H3C-135), 3,89/3.93 (H2C-10), 4.00 (HC-1), 5.34 (HcisC-32), 6.09 (HtransC-32), 6.43 (HC-31), 9.33 (HC-5), 9.51 (NH); 13C-NMR, (125 MHz, CD3OD, 26°C, tetramethylsilane) from heteronuclear single quantum correlation/heteronuclear multiple bond correlation spectra δ = 8.7 (71), 9.0 (121), 9.2 (181), 12.5 (21), 15.1 (82), 17.5 (81), 21.7 (171), 23.5 (10), 30.2 (20), 37.1 (15), 39.2 (172), 52.6 (135), 61.8 (1), 67.7 (132), 112.1 (12), 115.4 (18), 118.9 (32), 120.5 (17), 124.2 (19), 124.3 (16), 126.0 (8), 125.8 (13), 126.9 (31), 128.7 (3), 129.1 (6), 133.9 (11), 134.4 (7), 137.7 (9), 156.6 (2), 161.4 (14), 171.6 (133), 174.9 (4), 177.4 (5), 180.7 (173), 197.7 (131) (Figs. 2 and 4; and see Figs. 6–8, which are published as supporting information on the PNAS web site).
Fig. 2.
CD spectra of solutions in methanol of authentic Cj-NCC-2 (1, solid line) and the NCC 5a (dashed line), obtained from acid-induced isomerization of the FCC 2.
Preparation of the NCC 5a by Tautomerization of Ca-FCC-2 (2). A sample of 6 ml of the solution of Ca-FCC-2 (2) (16) [containing ≈42 nmol (26 μg) of 2] was first degassed, then acidified to pH 4.9 by addition of 0.6 ml of degassed buffer (1 M KH2PO4/K2HPO4) under an Ar atmosphere (pH control by pH electrode). The oxygen-free reaction mixture was stored at ambient temperature with protection from light for 20.5 h. The reaction mixture was then desalted by filtration through a Sep-Pak cartridge, followed by elution with methanol. The solution of the crude product in methanol exhibited the spectrum of an NCC, as analyzed by UV/visible spectroscopy, and was indicated to contain ≈46 nmol [29 μg, as calculated with log ε312nm (1) = 4.25] of 5a. The crude product obtained from this experiment, as well as the one of a second batch, in which 3 ml of the original solution of 2 was converted under the same conditions [giving 19 nmol (12 μg) of 5a], were purified with preparative HPLC [10 ml/min of 200 mM ammonium acetate in MeOH/water, 65:35 vol/vol (pH 6.8)]. The homogeneous fractions containing 5a (Rt = 21.9 min) were combined and desalted. The sample contained 38 nmol (24 μg) of the NCC 5a, according to its UV/visible absorbance spectrum.
Qualitative Kinetic Analysis of the Formation of the NCC 5a from the FCC 2. The preparative isomerization reaction was monitored by recording the UV/visible spectrum of the reaction mixture at 0, 15, 30, 90, 110, 270, and 1,230 min, and by analytical HPLC of samples of the reaction mixture collected at 0, 15, 110, 270, and 1,230 min (see Fig. 3). The observed time dependence of the relative amounts of 2, 5b, and 5a was analyzed based on a model of two consecutive first-order reactions (19), providing the estimated rates (at pH 4.9) of k1 = 0.025 min1 for the isomerization of 2 to 5b and k2 = 0.0055 min1 for the epimerization of 5b to 5a (for details see Fig. 9, which is published as supporting information on the PNAS web site).
Fig. 3.
Analysis of the acid-induced isomerization of the FCC 2. Progress of the reaction is monitored at specified times by analysis by HPLC (Upper, observation at 320 nm) and UV spectra (Lower).
Spectroscopic Characterization of the NCC 5a and Identification with Authentic Cj-NCC-2 (1). Spectroscopic data of 5a includes: MS (nano-electrospray ionization) m/z (%), 631.2 (12), 630.1 (52), 629.1 (100, [M+H]+), 597.1 (8, [M+H-32]+), 506.1 (5, [M+H-123]+); UV/visible, (MeOH) λmax (relative ε) 238 (sh, 1.00), 313 (0.62); CD, (MeOH) λmin/max (Δε) 226 (23.7), 252 (—8.6), 258 (—8.8), 281 (—18.4), 314 (5.1); 1H NMR (500 MHz, 0.22 mM in CD3OD, 26°C, tetramethylsilane), δ = 0.98 (H3C-82), 1.94 (H3C-21 and H3C-181), 2.07 (H3C-121), 2.24 (H3C-71), 2.28–2.36 (H2C-172), 2.42 (H2C-81), 2.46 (HAC-20), 2.61–2.67 (HAC-171), 2.72–2.78 (HBC-171), 2.84 (HBC-20), 3.74 (H3C-135), 3.92 (H2C-10), 3.94–3.99 (HC-1), 5.34 (HcisC-32), 6.08 (HtransC-32), 6.43 (HC-31), 9.33 (HC-5). The UV/visible, CD, 1H-NMR, and MS spectra of the NCC 5a showed no significant difference when compared with the spectra of authentic Cj-NCC-2 (1) (Figs. 2 and 4; and see Fig. 10, which is published as supporting information on the PNAS web site).
Fig. 4.
Section of 500-MHz 1H-NMR spectra (in deuteromethanol) of authentic Cj-NCC-2 (1, 0.34. and 0.07 mM, red traces) and the NCC 5a (0.22 mM, black trace), obtained from acid-induced isomerization of the FCC 2. The position of the signals of methylene protons (HBC-171 and HAC-20) is concentration dependent.
Identification of Cj-NCC-2 (1) and the NCC 5a by HPLC. Samples of the NCC 5a and authentic Cj-NCC-2 (1) from C. japonicum showed identical HPLC retention times and gave a single peak upon HPLC coinjection [flow: 500 μl/min; elution gradient at 0–40 min: from 200 mM ammonium acetate in MeOH/water 55:45 vol/vol (pH 6.8) to 100% 200 mM ammonium acetate in MeOH (pH 6.8); elution gradient at 40–60 min: 100% 200 mM ammonium acetate in MeOH (pH 6.8); see Fig. 11, which is published as supporting information on the PNAS web site].

Results

Analysis of the raw extracts of degreened leaves of C. japonicum by HPLC indicated the presence of Cj-NCC-1 (3) (as described earlier) (18), in addition to small amounts of a second, rather unpolar NCC, named Cj-NCC-2 (1) here. The two NCCs, 3 and 1, were found in a ratio of ≈23:1. Using repeated flash chromatography and reversed-phase HPLC, a pure sample of Cj-NCC-2 (1) could be isolated. The structure of authentic Cj-NCC-2 (1) was elucidated by spectroscopic means. The molecular formula of Cj-NCC-2 (1) was determined with high-resolution MS as C35H40N4O7, based on its pseudomolecular ion at m/z = 629.2990. This result indicated Cj-NCC-2 (1) has a molecular weight smaller than the one of any known NCC (6, 18, 2024) and relates to Cj-NCC-1 (3) (18) by its lack of one oxygen atom. The NCC 1 was thus identified as an isomer of a primary FCC, either the primary FCC (Bn-FCC-2) obtained from oil seed rape (Brassica napus) (13) or its C-1-epimer from sweet pepper (Capsicum annuum), named Ca-FCC-2 (2) (16). The UV/visible and CD spectra of a solution of 1 in methanol showed the absorption characteristics of the known NCCs (see Figs. 2 and 7; refs. 18 and 2024). Homonuclear and heteronuclear NMR spectroscopy provided consistent signal assignment for all 35 C atoms of 1 and for 34 H atoms, i.e., for all carbon-bound H atoms, except for the exchange labile HC-132. The spectroscopic data allowed the constitution of 1 to be determined as that of a 31,32-didehydro-1,4,5,10,15,20,22,24-octahydro-(132-methoxycarbonyl)-4,5-dioxo-4,5-secophytoporphyrin. The NCC 1 is a natural NCC lacking a peripheral oxidative refunctionalization. So far, a hydroxyl group or a polar, oxygen-bearing functional group at the ethyl group extending from ring B was considered to be a characteristic of the natural NCCs (6, 18, 2024). Such a function was assumed to be required as the basis for further peripheral refunctionalization, to assist the inter-organellar transport from the senescent chloroplast to the vacuole (17, 25, 26).
In earlier studies directed at determining the structures of FCCs, the FCC samples were noted to undergo uncharacterized chemical changes rather readily (13). Our interest has now turned to the question of whether FCCs would have a high tendency to engage in a specific tautomerization to NCCs, in analogy to the tautomerization reactions of hydroporphyrins, studied earlier by Eschenmoser (27). When a deoxygenated sample of ≈42 nmol (26 μg) of the chemically rather labile Ca-FCC-2 (2) (16) was stored at pH 4.9 in MeOH/water (30:70 vol/vol) at ambient temperature, a well-defined reaction was observed by UV/visible spectroscopy and HPLC. Conversion of the FCC 2 occurred within 120 min and a spectrum typical of an NCC appeared (see Fig. 3). The observed time dependence of the relative concentrations of 2, 5b, and 5a was analyzed by using a model of two consecutive reactions with first-order kinetics (see Fig. 9), and the isomerization of the FCC 2 to the NCC 5b could be estimated to occur with an effective rate of ≈0.025 min1 (half-life of 2 ≈27 min).
After overnight storage at ambient temperature, the crude product contained mainly the NCC 5a, according to HPLC analysis (Fig. 3). Monitoring by HPLC and UV/visible spectroscopy revealed the first product of the reaction not to be 5a, but a slightly less polar intermediate, the NCC 5b. Within ≈20 h the intermediate 5b disappeared nearly completely, with an estimated rate of ≈0.0055 min1 (i.e., half-life of 5b ≈130 min), to give the NCC 5a, which was isolated as the final isomerization product. The FCC 2 thus eventually isomerized in a rather clean (and nearly complete) reaction to the NCC 5a, which was isolated and subjected to spectroscopic and chromatographic analysis. The NCC 5a proved identical with Cj-NCC-2 (1). A sample of the tautomerization product 5a and a sample of authentic 1 were coinjected (at similar concentrations) for HPLC analysis, giving only one single HPLC peak (see Fig. 11). The electrospray ionization–MS, UV/visible, CD, and 1H-NMR spectral data collected for the sample of 38 nmol (24 μg) of NCC 5a were all indistinguishable from those obtained for the natural NCC, Cj-NCC-2 (1). The 1H-NMR spectra of 1 (in unbuffered solutions in methanol) showed an intriguing concentration dependence, resulting in slight chemical shift differences of the signals caused by H atoms of the propionic acid side chain mainly (see Figs. 4, 6, and 10). Only at comparable concentrations of 1 and 5a were the 1H-NMR spectra of the two samples practically identical. The observed concentration effects have not been examined further so far. They might be caused by a varying degree of dissociation of the acidic propionic acid function at C-173 or to (partial) intermolecular association (28).

Discussion

In senescing higher plants, Chl a and Chl b both are degraded within several days to pheophorbide a (57), which is rapidly broken down further to polar, colorless, and nonfluorescent Chl catabolites (see Fig. 1), which accumulate in the vacuoles of senescent leaves (5, 6, 15). Under conditions where senescenceinduced Chl breakdown occurred with high rates, FCCs were observed temporarily and suggested to be catabolic intermediates preceding the NCCs (29, 30).
The structures of two isomeric FCCs from two plant species have been analyzed so far (13, 16). Both were identified as primary FCCs, differing from their catabolic precursor, pheophorbide a, only by the minimal structural changes required by the enzymatic oxygenolytic ring opening reaction and the subsequent regioselective enzymatic reduction, which generates the C-1 stereocenter (13, 16). The two primary FCCs were found to be C-1 epimers (16), pointing to the existence of two classes of reductases in higher plants (12, 13) that introduce the stereocenter at C-1 stereoselectively but with either configuration. The reducing enzyme, tentatively named red Chl catabolite (RCC) reductase, was suggested to be inactive in a mutant of Arabidopsis thaliana, the leaves of which developed necrotic behavior (31).
The two epimeric primary FCCs set the stage for one of the two possible stereochemical lineages in Chl breakdown, which are both observed to a similar extent in higher plants (32). The absolute configuration at C-1 appears not to be altered during the subsequent catabolic transformations and the NCCs are also indicated to fall into the two groups, depending on the stereochemistry of their precursor FCC (22, 32): e.g., the NCCs 1 and 3 from leaves of C. japonicum fall into the same stereochemical group as the NCCs from C. annuum and are therefore derived from a primary FCC identical to Ca-FCC-2 (2) (16, 22). Despite the indicated stereochemical divergence of the NCCs at C-1, the long wavelength parts of the CD spectra of the known natural NCCs are all similar (see, e.g., Fig. 2; refs. 6, 18, and 2024). A common absolute configuration at the chiral centers at C-15 (and C-132) of the natural NCCs is thought to be responsible for the latter. The corresponding UV absorbance at ≈315 nm is caused by ring B (4, 5, 24), which experiences its nearest chiral center at C-15 (see Fig. 1 and ref. 33 for the numbering used). Analysis of the nuclear Overhauser effect spectra of various NCCs allowed the relative configuration at C-15 and C-132 to be derived as being the same in all of them. As configurational equilibration at the exchange labile C-132 produces the more stable trans arrangement of the bulky groups at C-15 and C-132, the configurationally stable chiral center C-15 directs the stereochemical preference at C-132 (4, 5, 18, 20, 2224). The absolute configuration of the stereocenters C-15 and C-132 of the NCCs is thus determined upon generation of the chiral center C-15.
We report here on an isomerization of the natural primary FCC 2 to the NCC 1, which occurs nonenzymatically and with high stereoselectivity at the crucial chiral center C-15. The NCC 1 (Cj-NCC-2) from extracts of senescent leaves of C. japonicum is a nonpolar natural NCC. The FCC 2 is identical to the unknown primary FCC from C. japonicum (22), the catabolic precursor of the NCC 1. The sample of 2 actually used here was obtained as Ca-FCC-2 from C. annuum (16).
In mildly acidic aqueous medium, at pH 4.9, the FCC 2 isomerizes to an NCC with a half-life of <30 min at ambient temperature. This behavior of the monoacid 2 contrasts with the much higher stability of the methyl esters of 2 and of isomeric FCCs (34). The remarkable effect of the type of carboxylic acid functionality provides direct evidence for a relevant role in the isomerization process of the propionic acid function extending from C-17, which is S-configurated in the FCC 2 but is prochiral in the NCCs. Indeed, as deduced from the mechanism inferred for the acid-catalyzed isomerization of the FCC 2 to the NCC 5a (= 1), the propionic acid function assists to accomplish the crucial step in the isomerization reaction (see Fig. 5) by stereoselective intramolecular protonation at C-15. Molecular models indicate such an intramolecular protonation to be feasible at C-15 from the re-side of the intermediate “enamine” 4 only, but not from the si-side, to result in R configuration at C-15 in 5a and 5b. Accordingly, the protonation product would not lead directly to 5a, an NCC with a stable relative trans arrangement of the bulky groups at C-15 and C-132. However, it would require the formation of an intermediate epi-NCC (5b) with the alternative cis arrangement. An unstable NCC is indeed observed as an intermediate in the isomerization reaction of the FCC 2 to the NCC 5a. This intermediary NCC is thus suggested to be (132epi)-5a (= 5b), the C-132 epimer of 5a.
Fig. 5.
Proposed mechanism of the acid-catalyzed isomerization of the FCC 2 to the NCC 5a (= 1). Tautomerization of 2 to the “enamine” 4 is suggested to occur by protonation at N-24, a relatively basic imino-position of the FCC, and loss of the proton HC-18. In weakly acidic medium the subsequent protonation of 4 at C-15 can be achieved in an intramolecular fashion by the propionic acid side chain at C-17. Intramolecular protonation by the propionic acid function is suggested to occur diastereoselectively from the re-side, generating the zwitter-ion iso-4. Deprotonation of iso-4 at C-17 yields 5b, the NCC, in which the methoxycarbonyl function at C-132 and ring D are cis to each other. The epimerization of 5b to the NCC 5a involves a β-keto-ester moiety and establishes a trans arrangement of the sterically demanding functions at C-132 and C-15 (with R configuration at C-15 and S configuration at C-132).
The deduced crucial role of the propionic acid function in the mechanism of the stereoselective isomerization of the FCC 2 to the NCC 1 (= 5a) provides a basis for a stereochemical assignment of C-15 and C-132 in the NCC 1. An intramolecular protonation involving a stereochemical transposition from C-17 to C-15 is invoked as the basis for the highly effective induction of the absolute configuration as 15-R (and 132-S). The absolute configuration at C-15 in 1 is the same as the one in the other known NCCs: two of the three chiral centers at the skeleton of the natural NCCs can now (tentatively) be assigned as 15-R and 132-S. The common stereochemistry of the NCCs at C-15 is a predictable outcome of the hypothetical nonenzymatic isomerization and is in striking contrast to the stereodivergence of the tetrapyrrolic Chl catabolites at C-1 (16), the chiral center introduced enzymatically during Chl breakdown. Such a surprising stereodivergence of an enzymatic step was rationalized as a feature of genuine catabolic processes (16), as it is likely to have no metabolic consequence. Considering this, the apparently common stereochemical build-up of the NCCs at their C-15 (and C132) is all the more paradoxical. As shown here, it can be rationalized as the consequence of an inherent tendency of the short-lived FCCs to isomerize spontaneously and stereoselectively in weakly acidic medium.
The structure of Cj-NCC-1 (3) and other polar natural NCCs point to refunctionalization reactions at several side chains (6, 18, 20, 23), which are likely to occur in the chloroplast and at the stage of the FCCs (15). Such peripheral transformations are indicated to assist the transport of the FCCs from the senescent chloroplast to the vacuole (17, 25, 26). According to the current topographical model of Chl breakdown (6, 8), FCCs are imported into the vacuoles before their transformation to the NCCs. The vacuoles, therefore, not only function as the place of storage for the NCCs in senescent plants (8, 14), but also as the site of their formation (5, 6).
Direct in vivo measurements of the vacuolar pH during senescence have not been performed. It is commonly accepted that the acidity of the vacuoles changes only marginally, if at all, as plant cells age. Thus, the rather acidic environment in the vacuoles (pH ≈4–5; refs. 17 and 35) would induce FCCs to undergo a rapid stereoselective isomerization to the NCCs, as shown here. The deduced half-life of the FCCs of 30 min or less (depending on the effective acidity) in the vacuole is kinetically competent with the temporary appearance of minor amounts of FCCs during senescence-induced Chl breakdown. Indeed, from comparison of the apparent rate of import of FCCs into the vacuole (limited by the overall rate of disappearance of pheophorbide a) with the rate of the isomerization of the FCC 2 at pH 4.9, FCCs can be estimated to represent only ≈1% (or less) of the total amount of Chl catabolites in the vacuoles.
The delineated kinetic and stereochemical criteria provide evidence for the relevance of a nonenzymatic isomerization of FCCs to NCCs during Chl breakdown. The FCCs are shown to be programmed by nature to further isomerize to the NCCs in an acidic medium. This feature of Chl breakdown has a remarkable manmade parallel in Woodward's total synthesis of Chl: There a surprisingly effective, stereoselective build-up of the ring D periphery of the chlorin macrocycle was achieved by strategic exploitation of the principle “that two things cannot be at the same place at the same time” (36, 37).
According to our studies the senescent leaf does not depend on an enzyme for the formation of NCCs from FCCs, an important step in Chl breakdown. Instead, interorganellar transport and a compartmentalized and suitable reaction medium (the acidic vacuolar sap) are used in an economic way to achieve this step in a well-controlled catabolic pathway. Export of the FCCs from nonacidic metabolically active compartments to the vacuole is recognized as a critical catabolic step in Chl breakdown (6, 8). In case this export to the vacuoles was slow, FCCs would accumulate in senescent leaves, as may upstream catabolites, like RCC. Indeed, accumulation of RCC is assumed to be responsible for the lesion mimic phenotype observed in A. thaliana mutants defective in RCC reductase (31).
The formation of NCCs was considered as the “ultimate” step in the breakdown of Chl in senescent higher plants (6). However, in the presence of air the NCCs are oxidized to rust-colored products (3, 4, 6). In some senescent leaves NCCs are processed further oxidatively to other tetrapyrroles (38) and Chl appears to be broken down even to monopyrrolic fragments (39). In view of the recent identification of the tetrapyrrolic heme degradation product bilirubin as an antioxidant cytoprotectant (40), the pronounced sensitivity of the NCCs to oxidants may turn out to have physiologic relevance in plant senescence.
In conclusion, the fleetingly observable FCCs can now be considered to be the proper products of the enzyme-catalyzed part of Chl breakdown in senescent higher plants. Once deposited in the vacuoles, the FCCs are geared to isomerize to NCCs, whose further fate and possible physiological role deserve further attention. These findings also underline the crucial position of the ring opening monooxygenase in aerobic Chl catabolism (10). Aside from being the only known senescence-specific enzyme in higher plants, this latter oxygenase or variants of it are also known to be the key enzymes in Chl catabolism in senescent green algae (6, 41).

Note

Abbreviations: Chl, chlorophyll; FCC, fluorescent Chl catabolite; NCC, nonfluorescent Chl catabolite; RCC, red Chl catabolite.

Acknowledgments

We thank Sonja Berger for technical assistance and Dr. Christian Eichmüller for measuring NMR spectra. This work was supported by Austrian National Science Foundation Projects 13503 and 16097, and Swiss National Science Foundation Grant 31.63628.

Supporting Information

HTML Page - 2207SuppText.html
HTML Page - 2207SuppText.html
HTML Page - 2207Fig10Legend.html
HTML Page - 2207Fig10Legend.html
Adobe PDF - 2207Fig10.pdf
Adobe PDF - 2207Fig10.pdf
HTML Page - 2207Fig11Legend.html
HTML Page - 2207Fig11Legend.html
Adobe PDF - 2207Fig11.pdf
Adobe PDF - 2207Fig11.pdf
HTML Page - 2207Fig6Legend.html
HTML Page - 2207Fig6Legend.html
Adobe PDF - 2207Fig6.pdf
Adobe PDF - 2207Fig6.pdf
HTML Page - 2207Fig7Legend.html
HTML Page - 2207Fig7Legend.html
Adobe PDF - 2207Fig7.pdf
Adobe PDF - 2207Fig7.pdf
HTML Page - 2207Fig8Legend.html
HTML Page - 2207Fig8Legend.html
Adobe PDF - 2207Fig8.pdf
Adobe PDF - 2207Fig8.pdf
HTML Page - 2207Fig9Legend.html
HTML Page - 2207Fig9Legend.html
Adobe PDF - 2207Fig9.pdf
Adobe PDF - 2207Fig9.pdf

References

1
Hendry, G. A. F., Houghton, J. D. & Brown, S. B. (1987) New Phytol. 107, 255–302.
2
Leeper, F. J. (1991) in Chlorophylls, ed. Scheer, H. (CRC, Boca Raton, FL), pp. 407–432.
3
Matile, P. (1987) Chimia 41, 376–381.
4
Kraütler, B., Jaun, B., Bortlik, K., Schellenberg, M. & Matile, P. (1991) Angew. Chem. Int. Ed. Engl. 30, 1315–1318.
5
Matile, P., Hörtensteiner, S., Thomas, H. & Kraütler, B. (1996) Plant Physiol. 112, 1403–1409.
6
Kraütler, B. & Matile, P. (1999) Acc. Chem. Res. 32, 35–43.
7
Rüdiger, W. (2002) Photosynth. Res. 74, 187–193.
8
Matile, P., Hörtensteiner, S. & Thomas, H. (1999) Annu. Rev. Plant Physiol. Plant Mol. Biol. 50, 67–95.
9
Hörtensteiner, S., Vicentini, F. & Matile, P. (1995) New Phytol. 129, 237–246.
10
Hörtensteiner, S., Wüthrich, K. L., Matile, P., Ongania, K.-H. & Kraütler, B. (1998) J. Biol. Chem. 273, 15335–15339.
11
Rodoni, S., Vicentini, F., Schellenberg, M., Matile, P. & Hörtensteiner, S. (1997) Plant Physiol. 115, 677–682.
12
Wüthrich, K. L., Bovet, L., Hunziker, P. E., Donnison, I. S. & Hörtensteiner, S. (2000) Plant J. 21, 189–198.
13
Mühlecker, W., Ongania, K.-H., Kraütler, B., Matile, P. & Hörtensteiner, S. (1997) Angew. Chem. Int. Ed. Engl. 36, 401–404.
14
Matile, P., Ginsburg, S., Schellenberg, M. & Thomas, H. (1988) Proc. Natl. Acad. Sci. USA 85, 9529–9532.
15
Hörtensteiner, S. & Kraütler, B. (2000) Photosynth. Res. 64, 137–146.
16
Mühlecker, W., Kraütler, B., Moser, D., Matile, P. & Hörtensteiner, S. (2000) Helv. Chim. Acta 83, 278–286.
17
Matile, P. (1997) Adv. Bot. Res. 25, 87–112.
18
Curty, C. & Engel, N. (1996) Phytochemistry 42, 1531–1536.
19
Frost, A. A. & Pearson, R. G. (1961) Kinetics and Mechanism (Wiley, New York).
20
Mühlecker, W. & Kraütler, B. (1996) Plant Physiol. Biochem. 34, 61–75.
21
Iturraspe, J., Moyano, N. & Frydman, B. (1995) J. Org. Chem. 60, 6664–6665.
22
Oberhuber, M., Berghold, J., Mühlecker, W., Hörtensteiner, S. & Kraütler, B. (2001) Helv. Chim. Acta 84, 2615–2627.
23
Berghold, J., Breuker, K., Oberhuber, M., Hörtensteiner, S. & Kraütler, B. (2002) Photosynth. Res. 74, 107–117.
24
Kraütler, B., Jaun, B., Amrein, W., Bortlik, K., Schellenberg, M. & Matile, P. (1992) Plant Physiol. Biochem. 30, 333–346.
25
Martinoia, E., Klein, M., Geisler, M., Sánchez-Fernández, R. & Rea, P. A. (2000) in Vacuolar Compartments, eds. Robinson, D. G. & Rogers, J. C. (Sheffield Academic, Sheffield, U.K.), Vol. 5, pp. 221–253.
26
Tommasini, R., Vogt, E., Fromenteau, M., Hörtensteiner, S., Matile, P., Amrhein, N. & Martinoia, E. (1998) Plant J. 13, 773–780.
27
Eschenmoser, A. (1988) Angew. Chem. Int. Ed. Engl. 27, 5–40.
28
Falk, H. (1989) The Chemistry of Linear Oligopyrroles and Bile Pigments (Springer, Vienna).
29
Bachmann, A., Fernández–López, J., Ginsburg, S., Bouwkamp, J. C., Solomos, T. & Matile, P. (1994) New Phytol. 126, 593–600.
30
Ginsburg, S., Schellenberg, M. & Matile, P. (1994) Plant Physiol. 105, 545–554.
31
Mach, J. M., Castillo, A. R., Hoogstraten, R. & Greenberg, T. (2001) Proc. Natl. Acad. Sci. USA 98, 771–776.
32
Hörtensteiner, S., Rodoni, S., Schellenberg, M., Vicentini, F., Nandi, O. I., Qui, Y.-L. & Matile, P. (2000) Plant Biol. 2, 63–67.
33
Scheer, H. (1991) in Chlorophylls, ed. Scheer, H. (CRC, Boca Raton, FL), pp. 1–30.
34
Oberhuber, M. & Kraütler, B. (2002) ChemBioChem 3, 104–107.
35
Hinder, B., Schellenberg, M., Rodoni, S., Ginsburg, S., Vogt, E., Martinoia, E., Matile, P. & Hörtensteiner, S. (1996) J. Biol. Chem. 271, 27233–27236.
36
Woodward, R. B., Ayer, W. A., Beaton, J. M., Bickelhaupt, F., Bonnett, R., Buchschacher, P., Closs, G. L., Dutler, H., Hannah, J., Hauck, F. P., et al. (1960) J. Am. Chem. Soc. 82, 3800–3802.
37
Woodward, R. B., Ayer, W. A., Beaton, J. M., Bickelhaupt, F., Bonnett, R., Buchschacher, P., Closs, G. L., Dutler, H., Hannah, J., Hauck, F. P., et al. (1990) Tetrahedron 46, 7599–7659.
38
Losey, F. G. & Engel, N. (2001) J. Biol. Chem. 276, 8643–8647.
39
Suzuki, Y. & Shioi, Y. (1999) Plant. Cell Physiol. 40, 909–915.
40
Baranano, D. E., Rao, M., Ferris, C. D. & Snyder, S. H. (2002) Proc. Natl. Acad. Sci. USA 99, 1693–1698.
41
Engel, N., Curty, C. & Gossauer, A. (1996) Plant Physiol. Biochem. 34, 77–83.

Information & Authors

Information

Published in

The cover image for PNAS Vol.100; No.12
Proceedings of the National Academy of Sciences
Vol. 100 | No. 12
June 10, 2003
PubMed: 12777622

Classifications

Submission history

Received: October 17, 2002
Published online: May 30, 2003
Published in issue: June 10, 2003

Acknowledgments

We thank Sonja Berger for technical assistance and Dr. Christian Eichmüller for measuring NMR spectra. This work was supported by Austrian National Science Foundation Projects 13503 and 16097, and Swiss National Science Foundation Grant 31.63628.

Authors

Affiliations

Michael Oberhuber
Institute of Organic Chemistry, University of Innsbruck, Innrain 52a, A-6020 Innsbruck, Austria; and Institute of Plant Sciences, University of Bern, CH-3013 Bern, Switzerland
Joachim Berghold
Institute of Organic Chemistry, University of Innsbruck, Innrain 52a, A-6020 Innsbruck, Austria; and Institute of Plant Sciences, University of Bern, CH-3013 Bern, Switzerland
Kathrin Breuker
Institute of Organic Chemistry, University of Innsbruck, Innrain 52a, A-6020 Innsbruck, Austria; and Institute of Plant Sciences, University of Bern, CH-3013 Bern, Switzerland
Stefan Hörtensteiner
Institute of Organic Chemistry, University of Innsbruck, Innrain 52a, A-6020 Innsbruck, Austria; and Institute of Plant Sciences, University of Bern, CH-3013 Bern, Switzerland
Bernhard Kraütler
Institute of Organic Chemistry, University of Innsbruck, Innrain 52a, A-6020 Innsbruck, Austria; and Institute of Plant Sciences, University of Bern, CH-3013 Bern, Switzerland

Notes

To whom correspondence should be addressed. E-mail: [email protected].
Communicated by Albert J. Eschenmoser, Swiss Federal Institute of Technology, Zürich, Switzerland, April 14, 2003

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Altmetrics




Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Breakdown of chlorophyll: A nonenzymatic reaction accounts for the formation of the colorless “nonfluorescent” chlorophyll catabolites
    Proceedings of the National Academy of Sciences
    • Vol. 100
    • No. 12
    • pp. 6893-7418

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media