You are on page 1of 11

Lysine catabolism in

Rhizoctonia leguminicola and


related fungi.
F P Guengerich and H P Broquist
J. Bacteriol. 1976, 126(1):338.

These include:
CONTENT ALERTS

Receive: RSS Feeds, eTOCs, free email


alerts (when new articles cite this article),
more

Information about commercial reprint orders: http://journals.asm.org/site/misc/reprints.xhtml


To subscribe to to another ASM Journal go to: http://journals.asm.org/site/subscriptions/

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

Updated information and services can be


found at:
http://jb.asm.org/content/126/1/338

Vol. 126, No. 1


Printed in U.SA.

JOURNAL OF BACTERIOLOGY, Apr. 1976, p. 338-347


Copyright 1976 American Society for Microbiology

Lysine Catabolism in Rhizoctonia legumirnicola


and Related Fungi
F. PETER GUENGERICH AND HARRY P. BROQUIST*
Department of Biochemistry, Vanderbilt University School of Medicine, Nashville, Tennessee 37232
Received for publication 3 December 1975

The various routes for the catabolism of ly- The possibility was considered that catabolism
sine have been reviewed by Rodwell (27). Al- via w-acetylated derivatives may constitute a
though the catabolic pathways have been well generalized degradative pathway of lysine medelineated for several groups of bacteria, the tabolism in those fungi synthesizing lysine via
catabolism of lysine in yeasts and fungi has not the homocitric-aminoadipic acid pathway (3).
been subjected to a great deal of study. The This paper reports a series of studies, both in
identification of E-N-acetyllysine, 8-acetami- vivo and in vitro, of the lysine catabolism of
dovaleric acid, a-hydroxy-e-acetamidohexanoic diverse fungi that lend support to this view.
acid, and glutaric acid as metabolites of L-ly- Figure 1 presents the relevant pathways consine in the yeast Hansenula saturnus led Roth- sidered. A brief account of this work has apstein to postulate a degradative pathway in- peared (F. P. Guengerich and H. P. Broquist,
volving these intermediates in yeasts (28, 31). Fed. Proc. 32:593, 1973).
The yeast Torula utilis accumulates 8-aminoMATERIALS AND METHODS
valeric acid and a-hydroxy-e-aminohexanoic
acid from a-amonoadipic acid, presumably via
Cultures. R. leguminicola was grown in stationlysine (32), whereas Mattoon and Haight (15) ary culture (Roux bottles) on 250 ml of red clover hay
described a lysine-requiring mutant of Saccha- extract medium as previously described (2, 10). All
romyces cerevisiae that accumulates glutaric other organisms were grown with rotary shaking at
C (20 ml of medium/shake flask). N. crassa lysine
acid. Schweet et al. (34) showed that in Neuro- 32
auxotrophs 15069 and 33933 were grown on the mespora crassa L-lysine is converted to L-pipecolic dium
of Horowitz and Beadle (11) containing 2.0 mM
acid and an N-acyl derivative of a-hydroxy-e- L-lysine.
(Mutant 15069 lacks saccharopine dehydroaminohexanoic acid. Meister et al. (18) later genase [33]; 33933 lacks cis-homoaconitase [1].) The
described a pyridine nucleotide-dependent re- yeasts S. cerevisiae S-228C (from F. Womack, Vanductase that reduces A'-piperideine-2-carbox- derbilt University) and H. saturnus (ATCC 2579)
ylic acid stereospecifically to L-pipecolic acid in were grown either on 1.4% Difco nitrogen base-1%
dextrose medium or as described by Rothstein (29),
the same organism.
Previous work from our laboratory with the respectively.
Whole cell incubations. Whole cell lysine incubafungus Rhizoctonia leguminicola has shown tions
were performed with N. crassa and the yeasts
that -E-N-acetyllysine and L-pipecolic acid ac- by
the appropriately labeled lysine (1 to 10
adding
cumulate from either D- or L-lysine; a lysine ,umol) to the
media (20 ml) of mature (5 day) culracemase was demonstrated (8). Moreover, D- tures; incubation proceeded at 32 C (with shaking)
lysine was metabolized to D-e-N-acetyllysine, for 12 to 48 h. N. crassa mycelia were drained,
whereas L-lysine gave rise to L-pipecolic acid rinsed, blended with 95% ethanol, and extracted in a
via A'-piperideine-2-carboxylic acid in vivo (8). Soxhlet apparatus overnight with the ethanolic fil338

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

The catabolism of lysine was studied in several yeasts and fungi. Results with
cell-free extracts of Rhizoctonia leguminicola support a proposed pathway involving (D- and L-) e-N-acetyllysine, a-keto-E-acetamidohexanoic acid, 8-acetamidovaleric acid, and 8-aminovaleric acid in the conversion of L-lysine to shortchain organic acids. Label from radioactive L-lysine was found to accumulate in
D- and L-E-N-acetyllysine, 8-acetamidovaleric acid, 8-aminovaleric acid, and
glutaric acid in cultures of R. leguminicola, Neurospora crassa, Saccharomyces
cerevisiae, and Hansenula saturnus, suggesting that the proposed w-acetyl
pathway of lysine catabolism is generalized among yeasts and fungi. In N.
crassa, as is the case in R. leguminicola, the major precursor of L-pipecolic acid
was the L-isomer of lysine; '"N experiments were consistent with Al'-piperideine2-carboxylic acid as an intermediate in the transformation.

LYSINE CATABOLISM IN FUNGI

VOL. 126, 1976

339

SACCHAROPINE
D-LYSINE

'

L-LYSINE

-KETO-E-NH2

,,AHEXANOATE

D-E-N-AcLYSINE *;o:-

L-C-N-AcLYSINE

c(-KETO-E-N-Ac HEXANOATE

- PIPERIDEINE 2- C 'ARBOXYLAT E

_ >-OH-e-N-Ac
HEXANOATE

L- PIPE CO LATE

i - N - Ac -VA LERAT E

oc-OH-E -NH2
H EXANOATE
\1-

S-NH2 -VALERATE --GLUTARIC


SEMIALDEHYDE

GLUTARATE ---ACETATE

FIG. 1. Postulated pathways of lysine catabolism in yeasts and fungi.


trate; the ethanolic extracts were concentrated in
vacuo. Whole yeast cultures were harvested by the
same procedure.
In some cases (see Table 2) radiotracers were
injected under the mycelium of mature (12 day) R.
leguminicola cultures into the medium (250 ml). In
most cases, the medium was drained off and the
mycelium was suspended in 50 ml of sterile water
containing the compound(s) of interest (unless otherwise noted, the level of addition of labeled lysine
was 1 to 10 jmmol/50 ml). Incubation proceeded at
room temperature (stationary culture) for 12 to 24 h.
The mycelia were harvested and worked up as in the
case of N. crassa described above.
Cell-free extracts and incubations. Cell-free extracts of R. leguminicola (10 to 12 days of age) were
prepared as previously described (8) and added to
test incubation systems on the basis of protein content (14); when noted, "desalting" of enzyme fractions was accomplished using a column (2.5 by 50
cm) of Sephadex G-25 (at 4 C). After incubation,
reactions were usually stopped by boiling (5 min) or
treatment with trichloroacetic acid or HCl04 and
centrifuged. Products and reactants were separated
by appropriate chromatographic procedures to permit quantitation of results after measurement of
radioactivity.
Chromatography. The ion-exchange system described by Piez et al. (25) was used to separate the
amino acids of concern, with the column modifications as previously described (8); the flow rate was
50 ml/h, and 5-ml fractions were collected. Acidic
compounds (void volume), 8-aminovaleric acid, E-Nacetyllysine, and pipecolic acid (order of elution)
were readily separated using this system (8). The
identities of all radioactive peaks from the columns
were confirmed by paper chromatography. Desalting of amino acids was achieved by applying sam-

ples to a column (1.2 by 25 cm) of Dowex 50 X 8 (H+


form, 200 to 400 mesh), eluting the citrate with 200
ml of distilled water, and eluting the amino acids
with 2 N NH40H, followed by in vacuo removal of
the ammonia.
Acidic metabolites were usually separated by
chromatography on Dowex 1 X 8 (28); the column
size wag 0.9 by 45 cm (200 to 400 mesh), and samples
of 1 to 2 ml in volume were applied at pH 8 to 9. The
identities of peaks of concern were confirmed in all
cases by paper and/or thin-layer chromatography
(TLC).
Alternatively, in several instances silicic acid
chromatography was used to separate acidic metabolites (see Fig. 2 and 3). Ethanolic cell extracts were
concentrated in vacuo to a small volume and applied
to a Dowex 50 X 8 column (1.2 by 25 cm, 200 to 400
mesh, H+ form). The acidic compounds were eluted
with 200 ml of distilled water and concentrated in
vacuo at 40 C to about a 1-ml volume. The samples
were acidified, triturated with silicic acid, and chromatographed by the procedure of Ramsey (26).
Paper chromatography systems employed Whatman no. 1 paper in a descending apparatus. The
solvent systems employed were: BAW (n-butanolacetic acid-water, 4:1:1); PAW (pyridine-acetic acidwater, 10:7:3); BPW (n-butanol-pyridine-water,
1:1:1). TLC systems utilized 0.1-, 0.25-, or 0.5-mm
layers of silica gel G; the solvent systems used were:
BMA (benzene-methanol-acetic acid, 8:1:1); EA
(ethanol-14% aqueous NH40H, 3:1); CMA (chloroform-methanol-14% aqueous NH40H, 2:2:1); CM
(chloroform-methanol, 1:1).
Assay of isotopes. Radioactive counting was as
previously described (8, 10); for double-label counting, the Packard 3320 spectrometer was set up such
that the 3H efficiency was 17% in the lower channel
and 0.1% in the upper channel. The '4C effilciency

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

I- co2

340

GUENGERICH AND BROQUIST

chromatography (BAW).
8-Acetamido-[5-'4C]valeric acid was prepared as
follows. Ethyl 8-acetamido-[5-'4C]valerate was synthesized according to Rothstein (29), starting with
9.80 g (250 ,Ci) of K'4CN (New England Nuclear
Corp.) and 2.5 g of ethyl 4-bromobutyrate (J. T.
Baker Co.). The ester was heated for 10 min in 20 ml
of 2 N NaOH at 100 C. After cooling, the solution
was lowered to pH 2 with 6 N HCI and extracted five
times with ethyl acetate. The combined organic layers were dried with anhydrous sodium sulfate and
reduced in vacuo. The residue was taken up in water, adjusted to pH 8.2, and applied to a column (45
by 1.2 cm) of Dowex 1 X 8 (acetate form, 200 to 400
mesh). The column was washed with 100 ml of distilled water; 8-acetamidovaleric acid was eluted
with 0.5 N acetic acid. The peak fractions were
combined and concentrated in vacuo; the product
migrated as a single peak on TLC (CMA, visualization with 12 vapor and with bromcresol green; all
radioactivity chromatographed with the product).

RESULTS
Conversion of lysine to pipecolic acid. Evidence has been presented that in R. leguminicola the nitrogen atom of L-pipecolic acid is
derived from the e-amino group of the L-isomer
of lysine, suggesting that A'-piperideine-2-carboxylic acid is an intermediate in the transformation (9). Because the latter compound has
been suggested as an intermediate in the formation of pipecolic acid in N. crassa (18, 34),
the work with R. leguminicola was repeated
with a lysine-requiring auxotroph of N. crassa.
N. crassa 15069 was grown for 5 days on
either 0.2 mM DL-E-'>NIlysine (44.8% excess
'"N) or 0.4 mM DL-[a-';N]lysine (45.6% excess
'-N). Gas-liquid chromatography-mass spectrometry (9) of the ethyl esters of the isolated
pipecolic acid fractions showed that the pipecolic acid of N. crassa 15069 grown on teI'Nilysine contained 35.0% excess '-N, whereas
that from the same auxotroph grown on [a'5N]lysine contained only 1.2% excess '-N.
Since the percentage of incorporation of lysine label into pipecolic acid was found to be
somewhat variable in N. crassa, the doublelabel technique of Leistner et al. (14) was used
to examine the stereospecificity of the transformation. The role of a lysine racemase in vivo
had been suggested by the experiments of
Schweet et al. (34), as lysine-requiring N.
crassa auxotrophs were found to grow on high
levels of D-lysine. Since the lysine racemases of
R. leguminicola and Pseudomonas putida are
strongly inhibited both in vitro and in vivo by
NH2OH (1, 19), this antimetabolite was added
in the N. crassa experiments. Table 1 shows
that in the control experiment (Column 1) the
expected ratio of unity (pipecolic acid ratio/
lysine ratio) was obtained within experimental
error. Experiments with _[1_-'4C]- and iLU'4Cilysines (Table 1) indicated that about 70%
of the pipecolic acid is derived from the Lisomer of lysine. The remaining 30% can be
explained by either (i) direct conversion of nlysine to pipecolic acid, or (ii) incomplete inhibition of a lysine racemase by NH,OH in these
experiments.
Lysine: e-N-acetyllysine interconversion.
The formation of E-N-acetyllysine from variously labeled lysines after incubation of whole
cultures of R. leguminicola has been previously
demonstrated (8). Using these same techniques, e-N-acetyllysine has been shown to accumulate from lysine in Neurospora and yeast
cultures, consistent with observations of earlier
investigators on the occurrence of e-acetyllysine derivatives in these organisms (28, 34).
Table 2 presents a series of such experiments in

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

was 50% in the upper channel and less than 10% in


the lower channel (corrections were made using
both [3H]- and [14C]toluene internal standards).
Incorporation of '5N into pipecolic acid was assayed as previously described (8).
Materials. D-, L-, and DL-e-N-acetyllysines were
purchased from Cyclo Chemical Corp.; 8-aminovaleric acid was from Calbiochem. Radioactive lysines
were purchased from New England Nuclear Corp.
[1,5-'4C]glutaric acid was from International Chemical and Nuclear Corp.
]3_[1-'4C]lysine was prepared as previously described (8); the optical purity was determined to be
99.0%. DL-[a-'3N]lysine and DL-[E-' N]lysine were
synthesized as previously described (8).
Acetyl coenzyme A was synthesized according to
Ochoa (23). 8-Aminovaleric acid was acetylated with
pyridine-acetic anhydride to form 8-acetamidovaleric acid according to Fieser (5), using the procedure
described for the acetylation of alanine.
DL-E-N-benzoyllysine (Cyclo Chemical Corp.) was
used to prepare DL-a-bromo-E-benzamidohexanoic
acid (9), which was used to synthesize DL-a-hydroxye-aminohexanoic acid according to Fischer and
Zemplen (6) (melting point, 199 to 204 C; calculated
for C6iH,,NO: C, 48.94; H, 8.91; N, 9.52; found: C,
48.85; H, 8.78; N, 9.46 [Galbraith Laboratories,
Knoxville, Tenn.]; single ninhydrin-positive spot on
paper chromatography [BAW]). Acetylation with
pyridine-acetic anhydride (5) followed by mild basic
hydrolysis produced DL-a-hydroxy-E-acetamidohexanoic acid.
Labeled E-N-acetyllysines were synthesized from
the appropriately labeled lysines as previously described (21). The products were purified by ion-exchange chromatography (25) and desalted prior to
use. a-Keto-E-acetamido-lG-3H]hexanoic acid was
prepared by treatment of L-E-N-acetyl-[G-3H]lysine
with Crotalus adamanteus venom (Ross Allen Reptile Institute, Silver Springs, Fla.) according to
Meister (16). The product was purified by chromatography on Dowex 1 X 8 (elution with distilled
water, 1 N HOAc, and 2 N HOAc in successive
order) followed by subsequent purification by paper

J. BACTERIOL.

LYSINE CATABOLISM IN FUNGI

VOL. 126, 1976

341

TABLE 1. Stereochemistry of the lysine (Lys) to pipecolic acid (Pip) conversion in N. crassaa
Labeled
Lysb
Labeled
Ly4C
3H
14C

(pLci)

DL-[4,5-3H]/DL-[2-'4C]

% Pip derived

Pip

Lys

(ACi) 3H/'4C

3H

(gCi) '4C (IMCi) 3H/1'4C

Pip ratio
ratio/Lys

from:

D-Lye' L-Lyed

0.064
1.07
0.280
4.39
(3.64)e (3.39)
0.052
8.89
1.42
35
65
0.463
6.28
18.75 2.99
DL-[4,5-3H/D[1- 14C]
(1.74)
(2.47)
71
0.70
29
0.458
0.086
5.34
7.58
20.50 2.71
DL-[4,5-3H]/L-[U-14C]
(3.17)
(2.23)
a N. crassa 15069 was grown at 32 C in three shake cultures on the basal media (25 ml/flask) of Horowitz
and Beadle (11) containing 1.0 jig of biotin per ml and 1.0 mM L-lysine. After 5 days the labeled lysine
mixtures (50 MuM total lysine) were added along with 0.4 mM carrier DL-lysine and 1.5 mM NH2OH. After 12
h at 32 C, pipecolic acid was isolated from the mycelial extracts using the described ion-exchange system
of Piez et al. (25), desalted, and counted as described.
b Prepared by mixing DL-[4,5-3H]lysine (250 uCi/umol), DL-[2-'4C]lysine (3.9 uCi/umol), D-[1-'4C]lysine
(1.96 MCi/,umol), and L-[U-'4C]lysine (38.5 MCi/,umol) to achieve the stated ratios of 3H/'4C.
' The percentage (y) of the product (pipecolic acid) derived from D-lysine was calculated using the
expression y = 50n/x, where n = 3H/'4C ratio of the substrate (DL-44,5-3H]lysine/D-[1-'4C]lysine) and x = 3H/
14C ratio of the product (pipecolic acid) (13).
d The percentage of pipecolic acid arising from L-lysine was calculated by difference.
e Values in parentheses are the percentage incorporated.

7.69

1.89

4.09

Organism

Added substrate
D (%)

L (%)

lated as e-NAcLys (%)

94
0.45
R. leguminicola
6
L-[U-'4C]Lys (0.4 MuM)
0.15
40
60
N. crassa 33933
L-[4,5-3HILys (10 MAM)
81
0.05
19
S. cerevisiae
L-[4,5-3H]Lys (10 MuM)
0.04
21
79
H. saturnus
L-[4,5-3H]Lys (10 MAM)
a
Mature cultures were incubated for 48 h with either L-[U-'4C]lysine (312 MCi/Mumol) or L-[4,5-3H]lysine
(250 MACi/Mumol). Ethanolic cell extracts were prepared as described and applied to a Dowex 50 column [Piez
et al. (25)] from which E-N-acetyllysine was isolated; its isomeric nature was determined by the D-amino acid
oxidase assay (8, 35).

which the configuration of the formed E-N-acetyllysine was determined after incubation of
the fungi and yeasts with labeled L-lysines.
Significant 1-isomer was present in all instances; the -isomer predominated in R. leguminicola, whereas the L-form was prominent
in the other organisms.
Extracts of R. leguminicola form e-N-acetyllysine from both and L-lysine (Table 3). E-Nacetyllysine formation required acetyl coenzyme A, but acetyl phosphate was not utilized.
(Hydroxylamine was included in the incubation mixture to minimize lysine racemase activity [8].) In other experiments (not shown), enzymic hydrolysis of e-N-acetyllysine to lysine
proceeded significantly faster (about 20-fold)
with the L- than the n-isomer. The enzymic
racemization of L-e-N-acetyllysine to the n-isomer was also demonstrated in R. leguminicola
extracts (using the n-amino acid oxidase assay
[35] in the presence of added pyridoxal phos-.
n-

phate). The significance of these findings in


relation to the stereospecific requirements for
lysine catabolism in R. leguminicola is considered more fully below.
Conversion of E-N-acetyllysine to 8-acetamidovaleric acid via a-keto-e-acetamido-hexanoic acid in R. leguminicola. An extract prepared from an R. leguminicola culture incubated with DL-[4,5-3H]lysine was chromatographed on a silicic acid column under conditions appropriate for separation of carboxylic
acids (26) (Fig. 2). The major peak was identified as 8-acetamidovaleric acid by TLC (BMA,
EA) and paper chromatography (BAW). The
percentage of the lysine label accumulated in
the 8-acetamidovaleric acid fraction was 0.9%.
The remainder of the radioactive eluates were a
series of carboxylic acids, which were not characterized. After acid hydrolysis (6 N HCl,
110 C, 12 h), the radioactivity of the presumed
8-acetamidovaleric acid chromatographed with

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

TABLE 2. Configuration of &N-acetyllysine (eN-AcLys) in several yeasts and fungia


Label accumue-N-AcLys

342

GUENGERICH AND BROQUIST

authentic 8-aminovaleric acid (TLC, CMA; paper chromatography, BAW). Labeled 8-acetamidovaleric acid was also identified by ionexchange chromatography on Dowex 2 X 8 (28)
in an ethanolic extract of a R. leguminicola
culture incubated with DL-[2-14C]lysine.
Since it seemed reasonable that 8-acetamidovaleric acid was arising from E-N-acetyllysine via decarboxylation of the a-keto acid of
the latter, desalted extracts of R. leguminicola
were incubated with D-e-N-acetyl-[4,5extracts

E-N-AcLygc

Systema

Substrateb

3H]lysine, a-ketoglutarate, and pyridoxal phosphate for 2 h at room temperature (conditions


as in Table 4). Chromatography of an acidic
ethyl acetate extract showed essentially one
radioactive peak (paper [BAW], Rf = 0.98; TLC
[BMAI, Rf = 0.79). Material in the ethyl acetate extract formed a 2,4-dinitrophenylhydrazone (paper [BAW], Rf = 0.91; TLC [BMA], Rf
= 0.62). Reduction of the latter with hydrogen
over PtO2 (17) yielded exclusively e-N-acetyllysine as demonstrated by paper chromatography
(BAW, Rf = 0.38) and TLC (BMA, Rf = 0.03),
thus establishing that the product of the enzyme reaction must be a-keto-E-acetamidohexanoic acid (all chromatography was done simul-

formed

(,umol)
0.52
0.30
0

Complete
i-Lysine
Complete
L-Lysine
Complete, cell extract
DL-Lysine
boiled
Less acetyl CoA
DL-Lysine
0.05
Less acetyl CoA plus 5
0.01
DL-Lysine
,umol of acetylphosphate
a
In addition to the substrate noted, the complete
incubation system contained: crude cell extract (5
mg), acetyl coenzyme A (CoA) (5 ,umol), NH2OH (7.2
Amol), and potassium phosphate (200 jumol, pH 7.5)
in 3.5 ml total volume.
b In each case, 5 ,umol of either D-[1-'4C]lysine
(0.40 ,Ci/umol), L-[U'4CIlysine (0.48 ,uCi/,umol), or
DL-[1-_4C]lysine (0.40 ,uCi/,umol) was added.
' After incubation for 2 h at 25 C, all reactions
were stopped by boiling for 5 min and centrifuged; eN-acetyllysine (E-N-AcLys) formation was measured
by counting radioactivity after separation by paper
chromatography (BAW).

TABLE 4. Transamination of e-N-acetyllysine in R.


leguminicola extracts

Systema

a-Keto-e-acetamidohexanoic acid

(,mol)
0.57
0.18
0.62
0.50
0.001
0.002

Complete ......................
Less a-KG ....................
Less pyruvate .................
Less PLP ......................
Less a-KG, pyruvate, PLP
Complete; enzyme extract boiled.
a Complete system: desalted R. leguminicola enzyme extract (2.5 mg), DL-e-N-acetyl-[4,5-3H]lysine
(8.4 ,umol; 0.32 uCi/Vumol), a-ketoglutarate (a-KG)
(15 ,mol), pyruvate (15 ,mmol), pyridoxal phosphate
(PLP) (0.4 ,umol), and potassium phosphate (165
,Lmol, pH 7.4) in 3.3 ml total volume. After incubation for 2 h at 25 C, reactions were stopped with 0.2
ml of 50% trichloroacetic acid and centrifuged; aketo-e-acetamido-hexanoic acid was determined in
the supernatants as described.

1000-J

0.

500-

X0. 00

ML
>
FIG. 2. Silicic acid chromatography of lysine catabolites of R. leguminicola. A mature R. leguminicola
culture was drained, suspended in 50 ml of 20 uM DL-t4,5-3H]lysine (200 paCiIp.mol), and incubated at room
temperature for 24 h. The ethanolic mycelial extract was treated as described and chromatographed on a
silicic acid column according to Ramsey (26). S-N-Ac-Val, &-Acetamidovaleric acid.

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

TABLE 3. Lysine acetylation in R. leguminicola

J. BACTERIOL.

LYSINE CATABOLISM IN FUNGI

VOL. 126, 1976

TABLE 5. Evidence for conversion of eN-acetyllysine


to 6-acetamidovaleric acid in R. leguminicola
extracts
8-Acetamidovaleric

acid (jAmol)
Substrateacd(.ol

D-e-N-acetyl-[2-'4C]lysine
146
(0.10
lCi/,umol; 29 umol)a
L-e-N-acetyl-[G-3H]lysine
13
(0.13 ,uCi/j,mol; 49 j,mol)P
a
The complete incubation system contained R.
leguminicola crude homogenate (10.0 mg), a-ketoglutarate (110 ,umol), thiamine pyrophosphate (2.8
,umol), pyridoxal phosphate (0.64 ,umol), (3-mercaptoethanol (1.5 ,umol), and potassium phosphate (300
,umol, pH 7.5) in 6.0 ml total volume. After incubation for 5 h at 25 C, the incubation was stopped with
0.2 ml of 50% trichloroacetic acid and centrifuged.
The supernatant was adjusted to pH 9 and chromatographed on a Dowex 1X8 column as described by
Rothstein (28); the 8-acetamidovaleric acid peak was
counted.
b The incubation was essentially as described in
footnote a, except that 6.4 mg of crude homogenate
was used.

leguminicola extract and thiamine pyrophosphate.


8-Acetamidovaleric acid catabolism in R.
leguminicola. Mature R. leguminicola cultures were incubated for 24 h with variously
labeled lysines; samples of the ethanolic extracts were chromatographed as before (8, 25).
The fractions corresponding to the peak between the 60- and 100-ml elution volumes were
desalted; the radioactivity in this peak was
shown to migrate with authentic 8-aminovaleric acid upon subsequent TLC (CMA) and paper chromatography (BAW). (Under the conditions, used, an artifact was routinely obtained
after in vacuo concentration [40 C] of the desalted fraction that could be extracted into chloroform at neutral pH and was identified [TLC,
CMA] as the lactam of 8-aminovaleric acid, 2piperidone.) In R. leguminicola cultures, it was
found that the addition of 1.2 mM NH2OH during incubation with lysine resulted in an enhanced accumulation of 8-aminovaleric acid, to
the extent of about 1% of the lysine label, as
opposed to 0.1 to 0.2% in the absence of NH2OH
(presumably pyridoxal-dependent transamination of 8-aminovaleric acid is inhibited). As
with 8-acetamidovaleric acid, label did not accumulate in 8-aminovaleric acid for - or DL-[1'4Cllysine; all other lysine labels used ([2-'4C],

[6-'4C], [4,5-3H]) were incorporated to approxi-

mately the same extent.


The deacetylation of 8-acetamidovaleric acid
to form 8-aminovaleric acid was demonstrated
to occur in R. leguminicola extracts. 8-Acetamido-[5-'4C]valeric acid (60.7 umol, 20.2 uCi/
mmol) was incubated with desalted R. leguminicola enzyme extract (5.0 mg) for 9 h at 25 C (30
mM potassium phosphate, pH 7.5; total volume, 6.2 ml). After boiling and centrifugation,
8-aminovaleric acid (1.31 ,umol) was identified
in the supernatant after ion-exchange chromatography (25). When the experiment was performed with boiled enzyme extract, only negligible 8-aminovaleric acid (0 to 0.1 ,umol) was
formed. Thus, all of the postulated enzymic
transformations between ilysine and 8-aminovaleric acid (Fig. 1) have been demonstrated
to occur in cell-free extracts of R. leguminicola.
A mature R. leguminicola culture was
drained, suspended on 8-acetamido-[5'4C]valeric acid, and incubated for 24 h; silicic
acid chromatography of the acidic and neutral
fraction (not bound to Dowex 50) revealed the
presence of essentially two radioactive peaks
(Fig. 3). The second was identified as residual
8-acetamidovaleric acid (TLC, EA, BMA; paper
chromatography, BAW). The first peak eluted
at the expected position for glutaric acid; the

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

taneously). Such transaminase activity was


studied in more detail, as outlined in Table 4.
a-Ketoglutarate was preferred over pyruvate
as the keto-acid; only slight if any dependence
on added pyridoxal phosphate was shown with
the desalted R. leguminicola extract.
In another experiment the transaminase system (as described in Table 4) was repeated, but
after 2 h thiamine pyrophosphate (4.3 umol)
and oxidized nicotinamide adenine dinucleotide
(13.8 umol) were added, and incubation was
continued for 2 h more. When the ethyl acetate
extract was chromatographed, virtually all of
the radioactivity was coincident with 8-acetamidovaleric acid (paper [BAW], Rf = 0.81; TLC
[BMA], Rf = 0.50) and distinct from a-keto-eacetamidohexanoic acid. The 1)-isomer of E-Nacetyllysine was found to be metabolized much
more rapidly to 8-acetamidovaleric acid than
was the L-isomer in a R. leguminicola cell-free
system (Table 5). The interpretation of these
results relative to the stereospecific requirements for lysine catabolism in R. leguminicola
is complicated, however, by the presence of a
deacylase and racemase for e-N-acetyllysine
and is considered below. In a similar system, it
was found (data not shown) that n-e-N-acetyl[1-'4C]lysine loses its carboxyl group as 14CO2 in
an enzymic reaction that is partially dependent
upon the presence of added thiamine pyrophosphate and pyridoxal phosphate. a-Keto-E-acetamido-[G-3H]hexanoic acid was converted to 8acetamidovaleric acid (28) in the presence of R.

343

344

GUENGERICH AND BROQUIST

J . BAcTERIOL.

1000-

0500a.

200

400

600

800

ML -*
FIG. 3. Silicic acid chromatography of b-acetamidovaleric acid catabolites of R. leguminicola. A mature R.
leguminicola culture was incubated with 30 mM b-acetamido-[5-'4C]valeric acid (202 Xilmmol) for 24 h at
room temperature. The ethanolic extract was treated as described and chromatographed on a silicic acid
column (26). GLUT, Glutaric acid; S-N-Ac VAL, S-acetamidovaleric acid.

identity was confirmed by TLC of the free acid


(BMA) and by gas chromatography (3% OV-17,
130 C) of the dimethyl ester (prepared by reflux
in methanol in the presence of dry HCl).
Incubation of a mature R. leguminicola culture with [1,5-'4C]glutaric acid and subsequent
silicic acid chromatography showed the conversion to a complex mixture of tricarboxylic acid
cycle and other carboxylic acids.
In vivo catabolism of lysine via w-acetylated
derivatives in N. crassa and S. cerevisiae.
Evidence for the accumulation of E-N-acetyllysine from i-lysine has been presented in Table
2. In a similar series of whole cell incubations
with mature N. crassa 33933 and S. cerevisiae
cultures, these organisms were also found to
accumulate radioactive 8-acetamidovaleric acid
after incubation for 24 h with DL-[4,5-3H]lysine
(incubation details as in Table 2; separation of
8-acetamidovaleric acid by ion-exchange chromatography on Dowex 1 X 8 [28]; identification
by paper chromatography [BAW]). As in the
case of R. leguminicola, the level of incorporation was about 1% from lysine; moreover, in
both cases, label from either D- or DL-L114C]lysine was not accumulated in 8-acetamidovaleric acid.
Mature cultures of N. crassa 33933 and 15069
and S. cerevisiae also accumulated label after
24 h from DL-12-'4C]-, DL-[6-'4C]-, and DL-[4,53H]lysine (but not DL-[1-_4C]- or ]-[1_-4C]lysine)
in 8-aminovaleric acid (incubation details similar to those in Table 2; separation as described
in the case ofR. leguminicola; identification in
all cases by paper chromatography [BAW,

BPW] and TLC [CMA]). The level of accumulation of label in 8-aminovaleric acid was quite
variable in these experiments, ranging from 1
to 5% of the proffered lysine label.
Radioactivity from i4G-3H]lysine was found
to accumulate in a compound identified as ahydroxy-E-acetamidohexanoic acid (separation
on Dowex 1 X 8 [28]) in mature N. crassa 33933
and S. cerevisiae cultures (24-h incubations; details similar to those of other lysine incubations; the extent of incorporation in both cases
was about 1%). The radioactivity of the fraction
chromatographed with authentic a-hydroxy-eacetamidohexanoic acid on TLC (EA, BMA)
and paper chromatography (BAW, BPW); after
acid hydrolysis (6 N HCl, 110 C, 12 h) the radioactivity was coincident with authentic a-hydroxy-e-aminohexanoic acid as judged by TLC
(CMA) and paper chromatography (BAW,
BPW).
DISCUSSION
The '-N experiments demonstrate that in N.
crassa 33933, as is the case in R. leguminicola,
the nitrogen atom of pipecolic acid is derived
from the E-nitrogen of lysine, consistent with
the role of Al-piperideine-2-carboxylic acid (but
not A1-piperideine-6-carboxylic acid) as an intermediate in the conversion of lysine to pipecolic acid in vivo.
The results shown in Table 1 also argue that
in N. crassa 33933, as in R. leguminicola (8),
pipecolic acid is preferentially derived from the
L-isomer of lysine. These findings with fungi
contrast with the work with bacteria (19), ani-

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

VOL. 126, 1976

345

of lysine in yeasts and fungi (12, 33). Presumably, this compound accumulated by reversal of
the action of saccharopine dehydrogenase, since
N. crassa 15069 (which lacks this enzyme) did
not accumulate saccharopine.
In contrast to other organisms (bacteria and
animals), no evidence for accumulation was observed for a-aminoadipic acid (4, 7, 30), glutamic acid (24), glutamine, piperidone-2-carboxylic acid (lactam of a-aminoadipic acid), or
8-aminovaleramide (27, 36) from lysine in the
yeasts and fungi studied here. The in vivo work
with the various yeasts and fungi supports the
basic scheme of lysine catabolism proposed for
H. saturnus (28). On the basis of our work and
that of Rothstein (28) and Schweet et al. (34),
we propose that the scheme of lysine catabolism
presented in Fig. 1 is peculiar to and generalized among yeasts and fungi, with certain limitations (yeasts do not appear to form pipecolic
acid; the a-hydroxy compounds do not appear to
be important in R. leguminicola; the arrows
leading from a-hydroxy-E-aminohexanoic acid
to glutarate are based on the findings of Rothstein [28] in H. saturnus). All of the proposed
steps between L-lysine and 8-aminovaleric acid
have now been demonstrated to occur in cellfree extracts of R. leguminicola.
It can be forcibly argued that D-lysine figures
prominently in lysine catabolism via -E-N-acetyllysine in R. leguminicola (Fig. 1). Thus (i) D
E-N-acetyllysine accumulates preferentially in
R. keguminicola (Table 2), (ii) the lysine racemase appears to favor formation of D-lysine
from the L-isomer (8), (iii) D-lysine appears to
be favored over L-lysine for formation of E-Nacetyllysine (Table 3), and (iv) D-E-N-acetyllysine is preferentially catabolized over the Lisomer in R. leguminicola extracts (Table 5).
But the inference that lysine catabolism in R.
leguminicola may proceed exclusively via D-EN-acetyllysine is hazardous at this time in view
of the activity of an L-E-N-acetyllysine deacylase and an E-N-acetyllysine racemase in these
crude extracts. Purification of E-N-acetyllysine-a-ketoglutarate transaminase and demonstration of its stereospecificity would shed
important light on this interesting problem. It
would be attractive, although perhaps too simplistic, to conclude that in the lysine metabolism of R. leguminicola, the L-isomer is concerned with anabolic reactions and that a racemase functions to convert the L-form to -lysine
when metabolic events in the cell dictate catabolism.
It is of interest that 8-aminovaleric acid is
also a key intermediate in lysine catabolism in
Pseudomonas (19, 27); however, our work

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

mals (7), and higher plants (13), in which pipecolic acid derives from D-lysine; the possibility
that pipecolic acid is derived mainly from 'lysine also in N. crassa has been proposed (20).
The experiments described in this paper utilized (i) a lysine auxotroph to block synthesis of
endogenous lysine, (ii) a racemase inhibitor,
and (iii) a double-label technique to minimize
interexperimental variations; although the
synthesis of as much as 30% of the pipecolic acid
from Dlysine cannot be ruled out, the data
argue strongly that most of the pipecolic acid
arises from L-lysine in N. crassa.
Label from DL-[R-3H]pipecolic acid (10) was
not incorporated into lysine in either R. leguminicola or N. crassa 33933; moreover, in R. leguminicola, N. crassa 33933, and S. cerevisiae
label from pipecolic acid was not incorporated
into the usual lysine catabolites deriving from
w-acetylation (vide supra), a-aminoadipic acid
(a pipecolic acid metabolite in animals [7, 301
and P. putida [4, 27]), or any significant levels
of acidic compounds. Of these organisms, only
R. leguminicola metabolizes pipecolic acid to
any significant extent; the novel alkaloids slaframine (10) and 3,4,5-trihydroxyoctahydro-1pyrindine (9) are derived from pipecolic acid.
(The yeasts S. cerevisiae and H. saturnus did
not accumulate pipecolic acid under the conditions used, in support of Rothstein's findings
[28] with H. saturnus.)
a-Hydroxy-E-acetamidohexanoic acid was
shown to accumulate from lysine in N. crassa
33933 and S. cerevisiae. Rothstein (28) has previously demonstrated such an accumulation in
H. saturnus; Schwett et al. (34) identified an Nacyl derivative of a-hydroxy-E-aminohexanoic
acid in N. crassa (the acyl moiety was not
identified). In our work, free a-hydroxy-E-aminohexanoic acid was also found to accumulate
from DL-[4,5-3H]lysine in H. saturnus, as previously demonstrated by Rothstein (28). Neither
compound was found to accumulate from lysine
in R. leguminicola in vivo; moreover, attempts
to convert a-keto-E-acetamidohexanoic acid to
the a-hydroxy compound in the presence of R.
leguminicola extracts and reduced nicotinamide adenine dinucleotide and/or reduced nictinamide adenine dinucleotide phosphate were
unsuccessful.
A radioactive peak arising from DL-[4,53H]lysine eluted from the Dowex 50 X 8 column
(8, 25) after 8-aminovaleric acid and before E-Nacetyllysine in the case of N. crassa 33933 but
with none of the other organisms. This compound was identified as saccharopine (paper
chromatography: BAW, BPW, water-saturated
phenol), the immediate biosynthetic precursor

LYSINE CATABOLISM IN FUNGI

346

GUENGERICH AND BROQUIST

ACKNOWLEDGMENTS
This work was supported by Public Health Service grant
AM-14338 from the National Institute of Arthritis, Metabolism and Digestive Diseases. F. P. G. was the recipient of
Public Health Service Predoctoral Fellowship GM-47480
from the National Institute of General Medical Sciences.
LITERATURE CITED
1. Broquist, H. P. 1971. cis-Homoaconitase, p. 114-118. In
D. McCormick and L. Wright (ed.), Methods in enzymology, vol. 17, part B. Academic Press Inc., New
York.
2. Broquist, H. P., and J. J. Snyder. 1971. Rhizoctonia
toxin, p. 319-333. In S. J. Ajl, S. Kadis, and T. C.
Montie (ed.), Microbial toxins, vol. 7. Academic Press
Inc., New York.
3. Broquist, H. P., and J. S. Trupin. 1966. Amino acid
metabolism. Annu. Rev. Biochem. 35:231-274.
4. Calvert, A. F., and V. W. Rodwell. 1966. Metabolism of
pipecolic acid in a Pseudomonas species. III. L-a-

Aminoadipate 6-semialdehyde: nicotinamide adenine


dinucleotide oxidoreductase. J. Biol. Chem. 241:409414.
5. Fieser, L. F. 1964. Experiments in organic chemistry,
p. 138. Heath, Boston, Mass.
6. Fischer, E., and G. Zemplen. 1909. Neue Synthese von
Amino-oxysauren und von Piperidon-Derivaten
Chem. Ber. 42:4878-4892.
7. Grove, J. A., T. J. Gilbertson, R. H. Hammerstedt, and
L. M. Henderson. 1969. The metabolism of D- and Llysine specifically labeled with '5N. Biochim. Bio-

phys. Acta 184:329-337.

8. Guengerich, F. P., and H. P. Broquist. 1973. Biosynthesis of slaframine, (1S,6S,8aS)-1-acetoxy-6-aminooctahydroindolizine, a parasympathomimetic alkaloid of fungal origin. II. The origin of pipecolic acid.
Biochemistry 12:4270-4274.
9. Guengerich, F. P., S. J. DiMari, and H. P. Broquist.
1973. Isolation and characterization of a I-pyrindine
fungal alkaloid. J. Am. Chem. Soc. 95:2055-2056.
10. Guengerich, F. P., J. J. Snyder, and H. P. Broquist.
1973. Biosynthesis of slaframine, (1S,6S,8aS)-1-acetoxy-6-aminooctahydroindolizine, a parasympathomimetic alkaloid of fungal origin. I. Pipecolic acid
and slaframine biogenesis. Biochemistry 12:42644269.
11. Horowitz, N. H., and G. W. Beadle. 1943. A microbio-

logical method for the determination of choline by use


of a mutant of Neurospora. J. Biol. Chem. 150:325333.
12. Kuo, M. H., P. P. Saunders, and H. P. Broquist. 1964.
Lysine biosynthesis in yeast: a new metabolite of aaminoadipic acid. J. Biol. Chem. 239:508-515.
13. Leistner, E., R. N. Gupta, and I. D. Spenser. 1973. A
general method for the determination of precursor
configuration in biosynthetic precursor-product relationships. Derivation of pipecolic acid from D-lysine,
and of piperidine alkaloids from L-lysine. J. Am.
Chem. Soc. 95:4040-4047.
14. Lowry, 0. H., N. J. Rosebrough, A. L. Farr, and R. J.
Randall. 1951. Protein measurement with the Folin
phenol reagent. J. Biol. Chem. 193:265-275.
15. Mattoon, J. R., and R. D. Haight. 1962. Glutaric acid
accumulation by a lysine-requiring yeast mutant. J.
Biol. Chem. 237:3486-3490.
16. Meister, A. 1954. The a-keto analogues of arginine,
ornithine, and lysine. J. Biol. Chem. 206:577-585.
17. Meister, A., and P. A. Abendschien. 1956. Chromatography of a-keto acid 2,4-dinitrophenylhydrazones and
their hydrogenation products. Anal. Chem. 28:171173.
18. Meister, A., A. N. Radhakrishnan, and S. D. Buckley.
1957. Enzymatic synthesis of L-pipecolic acid and Lproline. J. Biol. Chem. 229:789-800.
19. Miller, D. L., and V. W. Rodwell. 1971. Metabolism of
basic amino acids in Pseudomonas putida. Catabolism of lysine by cyclic and acyclic intermediates. J.
Biol. Chem. 246:2758-2764.
20. Muller, W.-U., and E. Leistner. 1975. Conversion of Dlysine to L-lysine via L-pipecolic acid in Neurospora
crassa. Z. Naturforsch. Teil C 30:253-262.
21. Neuberger, A., and F. Sanger. 1943. The availability of
the acetyl derivatives of lysine for growth. Biochem.
J. 37:515-518.
22. Numa, S., Y. Ishimura, T. Nakazawa, R. Okazaki, and
0. Hayaishi. 1964. Enzymatic studies on the metabolism of glutarate in Pseudomonas. J. Biol. Chem.
239:3915-3926.
23. Ochoa, S. 1955. Crystalline condensing enzyme from
pig heart, p. 685-694. In S. P. Colowick and N. 0.
Kaplan (ed.), Methods in enzymology, vol. 1. Academic Press Inc., New York.
24. Perfetti, R., R. J. Campbell, J. Titus, and R. A. Hart.
line. 1972. Catabolism of pipecolate to glutamate in
Pseudomonasputida. J. Biol. Chem. 247:4089-4095.
25. Piez, K. A., F. Irreverre, and H. L. Wolff. 1956. The
separation and determination of cyclic amino acids.
J. Biol. Chem. 223:687-697.
26. Ramsey, H. A. 1963. Separation of organic acids in
blood by partition chromatography. J. Dairy Sci.
46:480-483.
27. Rodwell, V. W. 1969. Carbon catabolism of amino acids,
p. 217-235. In D. M. Greenberg (ed.), Metabolic pathways, vol. 3. Academic Press Inc., New York.
28. Rothstein, M. 1965. Intermediates of lysine dissimilation in the yeast, Hansenula saturnus. Arch. Biochem. Biophys. 111:467-476.
29. Rothstein, M. 1954. The synthesis of 8-aminovaleric
acid-8-C 4. J. Am. Chem. Soc. 76:3038-3039.
30. Rothstein, M., K. E. Cooksey, and D. M. Greenberg.
1962. Metabolic conversion of pipecolic acid to a-aminoadipic acid. J. Biol. Chem. 237:2828-2830.
31. Rothstein, M., and J. L. Hart. 1964. Products of lysine
metabolism in yeast. Biochim. Biophys. Acta 93:439441.
32. Sagisaka, S., and K. Shimura. 1961. Studies in lysine
biosynthesis. II. Metabolic fate of DL-a-aminoadipiu
acid-6-C'4 in T. utilis. J. Biochem. (Tokyo) 49:392396.
33. Saunders, P. P., and H. P. Broquist. 1966. Saccharo.

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

strongly suggests that e-N-acetyllysine and 8acetamidovaleric acid (as opposed to &-aminovaleramide 127, 361) are intermediates in the
conversion of lysine to 8-aminovaleric acid. Presumably, the degradation of 8-aminovaleric
acid via glutarate follows steps resembling
those delineated in Pseudomonas by Numa et
al. (22).
The evidence presented for a generalized
pathway of lysine catabolism specific for yeasts
and fungi is of evolutionary significance, as
these organisms (and only these organisms)
synthesize lysine via the previously described
homocitric-aminoadipic acid pathway (3).
Thus, these lower plants may be evolutionarily
related not pnly with respect to their biosynthetic enzymes of lysine metabolism but with
respect to their major catabolic enzymes as
well.

J. BACTERIOL.

VOL. 126, 1976


pine, an intermediate of the aminoadipic acid pathway of lysine biosynthesis. IV. Saccharopine dehydrogenase. J. Biol. Chem. 241:3435-3440.
34. Schweet, R. S., J. T. Holden, and P. H. Lowy. 1954.
The metabolism of lysine in Neurospora. J. Biol.
Chem. 211:517-529.

LYSINE CATABOLISM IN FUNGI

347

35. Soda, K., and T. Osumi. 1971. Amino acid racemase


(Pseudomonas striata), p. 629-636. In D. McCormick
and L. Wright (ed.), Methods in enzymology, vol. 17,
part B. Academic Press Inc., New York.
36. Takeda, H., and 0. Hayaishi. 1966. Crystalline L-lysine
oxygenase. J. Biol. Chem. 241:2733-2736.

Downloaded from http://jb.asm.org/ on February 22, 2014 by guest

You might also like