You are on page 1of 16

Subscriber access provided by LIBRARY OF CHINESE ACAD SCI

Perspective
Evidence for the Likely Origin of Homochirality in Amino
Acids, Sugars, and Nucleosides on Prebiotic Earth
Ronald Breslow
J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/ja3012897 • Publication Date (Web): 25 Mar 2012
Downloaded from http://pubs.acs.org on April 12, 2012

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 15 Journal of the American Chemical Society

1
2
3
4
Evidence
for
the
Likely
Origin
of
Homochirality
in
Amino
Acids,
Sugars,
and

5 Nucleosides
on
Prebiotic
Earth

6 

7 Ronald
Breslow

8 Department
of
Chemistry,
Columbia
University

9
10
New
York
NY
10024

11 rb33@columbia.edu

12 

13 Abstract.




14 Over
the
past
century
the
origin
of
terrestrial
prebiotic
homochirality
has
been
the

15
16
subject
of
many
speculations.

For
life
to
start
on
earth
and
elsewhere,
it
is
critical

17 that
the
building
blocks
of
amino
acids,
sugars,
and
nucleosides
be
created
in

18 predominant
homochiral
form.

Recent
findings
of
a
modest
excess
L
chirality
of
α‐
19 methylamino
acids
in
some
meteorites
that
landed
on
earth
have
furnished
an

20
important
piece
of
evidence.

We
have
shown
how
these
meteoritic
components
can

21
22 furnish
normal
L
amino
acids,
and
therefrom
D
sugars
and
nucleosides,
in
high

23 chiral
excess
under
sensible
prebiotic
conditions.

Some
important
remaining
goals

24 are
also
described.

25 

26
27

 Introduction.

28 Currently
 we
 are
 not
 surprised
 that
 L
 amino
 acids
 and
 D
 sugars
 are
 produced
 in

29 biological
systems,
since
the
enzymes
that
produce
them
are
themselves
homochiral

30 (not
 a
 mixture
 with
 their
 mirror
 images.)
 
 On
 prebiotic
 earth
 no
 such
 chirally

31 selective
catalysts
were
there
to
make
the
first
amino
acids
or
sugars
or
nucleosides,

32
33
so
 many
 scientists
 have
 speculated
 on
 how
 such
 selectivity
 could
 have
 arisen
 in
 a

34 previously
 achiral
 world.
 
 Some
 have
 invoked
 chiral
 faces
 of
 quartz,
 some
 have

35 suggested
 an
 accident
 that
 was
 then
 promulgated,
 but
 the
 question
 has
 had
 no

36 important
new
impetus
until
recently.
In
1969
a
carbonaceous
chondritic
meteorite

37 landed
in
Murchison
Australia
carrying
many
organic
compounds
(see
an
extensive

38
39
recent
review
by
Pizzarello
and
Groy1
including
the
earliest
work
by
Kvenholden
et

40 al.2
 and
 by
 Cronin
 and
 Pizzarello3
 cited
 in
 the
 review.)
 These
 compounds
 were

41 apparently
able
to
survive
the
frictional
heating
as
the
meteorite
passed
through
our

42 atmosphere
since
they
were
initially
at
ca.
10K,
and
chondritic
meteorites
are
pieces

43 of
 rock,
 with
 low
 thermal
 conductivity,
 from
 the
 asteroid
 belts
 that
 surround
 the

44
45 sun.

When
the
meteorite
was
split
open
the
interior
was
still
cold
enough
to
freeze

46 water.



47 
 

48 Among
the
compounds
identified
were
the
amino
acids
alanine,
valine,
aspartic
acid,

49 glutamic
acid,
proline,
and
leucine,
which
were
racemic,
with
equal
mixtures
of
the
L

50
51 and
D
forms,
along
with
achiral
glycine.
However,
five
amino
acids
were
found
that

52 had
 methyl
 groups
 instead
 of
 hydrogens
 on
 their
 alpha
 positions
 (Figure
 1),
 and

53 these
had
a
range
of
small
excesses
of
the
enantiomers
originally
described
as
the
L

54 amino
acids
(in
modern
terminology
they
are
the
S
enantiomers).

Since
that
time,

55
these
and
other
α‐methyl
amino
acids
with
small
excesses
of
the
S
enantiomer
have

56
57 been
found
in
the
Murchison,
Murray,
and
Orgueil
meteorites.1

58
59
60

 1

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 2 of 15

1
2
3
4













5 + + +
H 2N H3N CO2- H3N CO2-
6 CO2-
7 Me Pr Me Me Et
8
9
10 e.e. 2.8% S e.e. 2.8% S e.e. 15.2% S
11
12 + +
H 3N H 3N CO2-
13 CO2-
14 Me Me Me
Me (3R)
15 (3S)
Et Et
16
17 e.e. 9.1% S
18
e.e. 7% S 

19 

20 Figure
 1.
 
 Some
 α‐methyl
 amino
 acids
 discovered
 in
 the
 Murchison
 and
 other

21 meteorites,
 all
 of
 which
 have
 a
 small
 excess
 of
 the
 S
 configuration
 that
 was

22 described
 as
 L.
 
 The
 enantioexcesses
 showed
 a
 range
 of
 values
 in
 this
 and
 later

23
24
work.

25 

26 How
do
we
know
that
they
were
not
simply
contaminants
from
earth,
added
after

27 the
 meteorites
 landed?
 
 
 There
 are
 three
 arguments
 against
 this.
 
 First
 of
 all,
 the

28 actual
α‐methyl
amino
acids
isolated
from
the
Murchison
meteorite
are
not
found
in

29
30 our
 biology
 today,
 but
 the
 meteorite
 landed
 here
 only
 41
 years
 ago.
 
 Against
 this,

31 there
are
some
α‐methyl
amino
acids
produced
by
microorganisms4‐6
but
not
all
the

32 ones
in
the
meteorites.



Secondly,
enzymes
do
not
produce
compounds
with
only

33 partial
chiral
selectivities;
such
partial
chiralities
are
really
symptomatic
of
species

34
35
that
 were
 originally
 racemic
 but
 have
 then
 been
 partially
 deracemized.
 
 If

36 microorganisms
 had
 invaded
 the
 meteorites
 after
 they
 landed,
 they
 could
 perhaps

37 selectively
 cause
 destruction
 of
 the
 
 D
 enantiomers
 of
 the
 amino
 acids
 if
 they
 had

38 originally
been
racemic,
but
there
is
no
evidence
for
such
an
unlikely
process.

I
will

39 describe
an
alternative
scenario
for
such
partial
deracemization
below.

Finally,
the

40
41
α‐methyl
 amino
 acids
 have
 levels
 of
 13C
 and
 non‐exchangeable
 deuterium
 much

42 higher
 than
 are
 seen
 in
 molecules
 formed
 on
 Earth.
 
 These
 high
 levels
 of
 heavy

43 isotopes
 are
 generally
 seen
 in
 atoms
 delivered
 from
 space,
 where
 isotopic

44 fractionation
is
performed
at
very
low
temperatures.

They
are
generally
accepted
to

45
be
definitive
proof
that
the
species
examined
are
not
of
terrestrial
origin.

46
47 
 

48 Microwave
spectroscopy
has
detected
hundreds
of
molecules
in
the
interstellar
gas

49 clouds
in
space,
and
they
include
ammonia,
hydrogen
cyanide,
hydronium
ion,
and

50 various
ketones
and
aldehydes
and
the
acetylenes
from
which
they
can
be
derived.7

51
Thus
 the
 meteoritic
 unmethylated
 and
 methylated
 amino
 acids
 were
 probably

52
53 formed
 by
 Strecker
 reactions8
 from
 such
 species.
 The
 reactants
 are
 aldehydes
 for

54 normal
unmethylated
amino
acids
and
methyl
ketones
for
the
α‐methyl
amino
acids,

55 along
with
HCN,
NH3,
and
H3O+.

Methyl
ketones
are
the
products
from
reaction
of

56 terminal
acetylenes
with
hydronium
ion.

The
multimolecular
reactions
would
occur

57
58 on
 solid
 particles
 whose
 larger
 versions
 are
 asteroids,
 and
 they
 would
 be
 initially

59
60

 2

ACS Paragon Plus Environment
Page 3 of 15 Journal of the American Chemical Society

1
2
3
4
formed
 as
 racemates.
 
 They
 could
 develop
 some
 excess
 of
 the
 L
 forms
 if
 the

5 racemates
 absorbed
 circularly
 polarized
 light
 that
 selectively
 destroyed
 one

6 enantiomer,
perhaps
also
using
another
process
to
amplify
them.9


7 
 

8 William
 Bonner
 had
 shown
 that
 right
 circularly
 polarized
 light
 at
 uv
 wavelengths

9
10
would
 selectively
 destroy
 the
 D
 component
 of
 racemic
 leucine,
 leaving
 a
 small

11 excess
of
the
L
amino
acid.10

Astronomers
have
detected
a
small
excess
of
circularly

12 polarized
 infrared
 light
 in
 space,
 whose
 energy
 is
 too
 low
 to
 deracemize
 amino

13 acids.11‐13
However,
one
author
indicated
that
the
same
processes
that
produce
the

14 identified
 circularly
 polarized
 light
 in
 the
 infrared
 might
 also
 produce
 it
 in
 the

15
16
ultraviolet.12
 Very
 high
 energy
 light
 cannot
 penetrate
 our
 atmosphere
 for

17 observation
as
infrared
can,
and
could
not
penetrate
it
when
our
early
atmosphere

18 consisted
 of
 nitrogen
 and
 CO2.
 
 As
 one
 possibility,
 favored
 by
 many
 astronomers,

19 circularly
polarized
light
could
be
formed
in
the
universe
by
cyclotron
processes
in

20 neutron
 stars,
 just
 as
 experimental
 cyclotrons
 produce
 circularly
 polarized
 light

21
22 with
 opposite
 chirality
 above
 and
 below
 the
 circulation
 plane.
 
 Some
 other

23 astronomers
 prefer
 that
 circularly
 polarized
 light
 could
 be
 produced
 by
 magnetic

24 dwarf
stars.
One
of
the
important
challenges
for
astronomy
is
to
observe
outside
our

25 atmosphere
and
detect
circularly
polarized
short‐wavelength
light,
if
it
is
there.

For

26 instance,
 observations
 could
 be
 made
 from
 orbiting
 satellites,
 from
 the
 Hubble

27
28 telescope,
 or
 from
 the
 new
 Web
 Space
 Telescope.
 
 I
 am
 trying
 to
 get
 such

29 observations
made.



30 
 

31 If
there
was
also
(yet
undetected)
right
circularly
polarized
light
with
energy
in
the

32
uv
or
higher
irradiating
the
asteroid
belt
when
the
amino
acids
were
present
on
a

33
34 particle
that
later
came
to
Earth,
this
could
account
for
the
small
excesses
of
the
L

35 enantiomers
 seen
 in
 the
 α‐methyl
 amino
 acids.
 
 If
 this
 also
 happened
 with
 the

36 unmethylated
 amino
 acids
 that
 are
 found
 in
 our
 proteins
 they
 could
 racemize
 by

37 reversible
 loss
 of
 their
 alpha
 protons
 over
 time,
 explaining
 why
 they
 are
 found
 in

38
39
racemic
form,
but
no
such
racemization
is
possible
with
the
α‐methyl
amino
acids.

40 (Isoleucine
has
two
chiral
centers,
and
it
was
not
completely
racemic.


The
α‐center

41 would
 be
 equilibrated
 by
 reversible
 loss
 of
 its
 somewhat
 acidic
 proton
 but
 the
 β‐
42 center
 could
 not
 be
 equilibrated,
 so
 isoleucine
 could
 not
 be
 racemized
 by
 half

43
44 conversion
to
its
mirror
image.)





45 
 

46 Thus
an
attractive
idea
is
that
the
L
amino
acids
and
the
D
sugars
that
are
now
the

47 basis
of
life
were
first
“seeded”
by
the
arrival
on
Earth
of
α‐methyl
amino
acids
in

48
49
meteorites,
 formed
 on
 asteroid
 fragments
 as
 racemates
 and
 then
 partially

50 deracemized
by
circularly
polarized
light
of
short
wavelength.14
Light
of
the
needed

51 wavelength
 could
 not
 penetrate
 Earth’s
 atmosphere—now
 or
 in
 the
 past—so
 the

52 deracemization
occurred
in
space
and
the
result
was
then
delivered
to
Earth.

Such

53 homochirality
 was
 probably
 important
 since
 mixture
 of
 enantiomers
 would
 not

54
55
form
 well‐defined
 structures
 on
 random
 incorporation
 into
 polypeptides
 or

56 polysaccharides
 or
 polynucleotides.
 
 In
 fact,
 it
 seems
 likely
 that
 homochirality
 is

57 generally
necessary
for
the
origin
of
life
anywhere,
so
without
the
kind
of
processes

58
59
60

 3

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 4 of 15

1
2
3
4
that
 started
 it
 on
 earth
 other
 planets
 would
 not
 produce
 life.
 
 On
 Earth
 the
 amino

5 acid
 chirality
 was
 what
 we
 call
 L,
 but
 on
 other
 planets
 the
 mirror
 image
 systems

6 could
occur
by
the
processes
we
describe.



7 
 

8 We
 have
 recently
 shown
 how
 such
 α‐methyl
 amino
 acids
 can
 generate
 normal

9
10 unmethylated
 biological
 amino
 acids
 with
 some
 chirality
 transfer,
 and
 how
 that

11 excess
 chirality
 can
 be
 amplified
 under
 credible
 prebiotic
 conditions,
 producing

12 aqueous
solutions
with
very
high
enantioexcesses.

We
have
also
shown
how
the
D

13 sugars
and
D
ribonucleosides
could
have
arisen
on
prebiotic
Earth.


14

15
16 
 Results
and
Discussion

17
18 We
 addressed
 two
 questions.
 
 1)
 What
 credible
 prebiotic
 chemistry
 could
 use
 the

19 small
 excess
 S
 (L)
 chirality
 of
 the
α‐methyl
 amino
 acids
 in
 a
 meteorite
 to
 seed
 the

20
21
formation
 of
 L
 amino
 acids
 without
 such
 methyl
 groups?
 2)
 How
 could
 credible

22 prebiotic
processes
amplify
small
enantioexcesses
to
available
high
enantiopurity
in

23 solution?

The
ideal
is
of
course
100%,
but
enantioexcesses
of
90%
or
better
could

24 probably
be
enough
for
primitive
biology
to
select
the
dominant
enantiomer.

25 



26
27

 Formation
of
non­methylated
amino
acids
with
some
L
excess.



28 We
 devised
 a
 process
 of
 decarboxylative
 transamination:
 α‐methyl
 amino
 acids

29 reacted
 with
 α‐keto
 acids
 to
 form
 unmethylated
 amino
 acids
 on
 simple
 heating
 in

30 the
 presence
 of
 Cu(II)
 (Figure
 2).15
 
 This
 was
 based
 on
 earlier
 work
 we
 had
 done

31
32
showing
 that
 α‐methyl
 amino
 acids
 could
 perform
 such
 decarboxylative

33 transaminations
with
pyridoxal.16

34 


35 We
 observed
 some
 chiral
 transfer,
 in
 which
 the
 reaction
 of
 100%
 e.e.
 S‐α‐
36
37
methylvaline
 with
 sodium
 phenylpyruvate
 led
 to
 a
 37%
 enantioexcess
 of
 L‐
38 phenylalanine,
 and
 the
 same
 process
 with
 sodium
 pyruvate
 led
 to
 a
 20%

39 enantioexcess
 of
 L‐alanine.
 
 The
 reaction
 of
 sodium
 dimethylpyruvate
 with
 S‐α‐
40 methylisoleucine
 afforded
 L‐valine
 with
 17%
 transfer
 of
 chirality.
 
 All
 these

41
processes
 were
 successful
 only
 in
 the
 presence
 of
 catalytic
 Cu(II),
 a
 component
 of

42
43 some
 meteorites
 and
 surely
 present
 on
 prebiotic
 earth.
 
 Without
 it,
 or
 with
 Zn(II)

44 instead,
the
reactions
were
much
less
successful,
and
the
correct
chiral
transfer
was

45 not
observed.


46 

47
48
49
50
51
52
53
54
55
56
57
58
59
60

 4

ACS Paragon Plus Environment
Page 5 of 15 Journal of the American Chemical Society

1
2
3 R
4 R CO2 -
+ - H2O
5 H 3N N - CO2
6 O- CO2- CO2-
O +
7 Me Me Cu(II)
O + H2O
8
9
10 R R H
CO2 - + H+ CO2 -
11 H2O R CO2 -
N N +
12
H NH3 +
13 Me Me O Me
14 

15 

16
17
18 Figure
2.

Transaminative
decarboxylation
of
the
imine
from
S‐α‐methylvaline
and

19 an
α
ketoacid.

In
the
presence
of
Cu(II),
which
catalyzes
the
reaction,
the
product

20 amino
acid
has
an
excess
of
the
L
enantiomer,
but
not
without
the
Cu(II).

This
is

21
22
expected
if
a
square
planar
complex
is
formed
between
two
imines
and
the
Cu(II),
so


23 after
decarboxylation
of
one
ligand
it
is
protonated
selectively,
guided
by
the

24 stereochemistry
of
the
other
ligand.


25 

26 The
 process
 was
 credible
 under
 prebiotic
 conditions,
 involving
 only
 heating
 the

27
28
components
 in
 the
 presence
 of
 some
 catalytic
 Cu(II)
 ion.
 
 As
 we
 described,15
 DFT

29 calculations
indicated
that
a
square
planar
Cu(II)
complex
of
two
imines
that
were

30 formed
 from
 the
 amino
 acid
 and
 the
 keto
 acid
 would
 decarboxylate
 one
 of
 them,

31 which
would
then
protonate
to
form
the
L
amino
acid
steered
by
the
chirality
of
the

32 second
 ligand,
 as
 we
 observed.
 
 However,
 the
 few
 percent
 of
 S
 excess
 in
 the

33
34 meteoritic
amino
acids
would
produce
only
a
1%
or
so
enantioexcess
in
the
product

35 amino
acids.

Amplification
of
the
resulting
chiral
excess
would
be
critical
to
achieve

36 a
 dominance
 of
 the
 L
 amino
 acids
 large
 enough—of
 the
 order
 of
 a
 90/10
 ratio
 or

37 better
 of
 L
 to
 D—that
 new
 organisms
 or
 pre‐organisms
 would
 select
 them
 for

38 evolutionary
success.

39
40 
 Amplification
of
the
L­amino
acid
excesses
in
solution.


41 In
 1969
 Morowitz
 had
 proposed
 a
 process
 by
 which
 a
 small
 enantioexcess
 of
 an

42 amino
 acid
 could
 be
 amplified
 to
 form
 solutions
 with
 large
 enantioexcesses
 under

43 simple
prebiotic
conditions.17
This
was
an
amplification
of
the
concentration
of
the

44 excess
component
in
solution
by
concentrating
it
and
removing
the
racemate,
not
by

45
46 making
 more
 of
 it.
 
 We
 conceived
 the
 same
 idea
 in
 2006,18
 not
 aware
 of
 his
 work,

47 and
Blackmond
also
published
a
version
of
the
concept
in
2006
slightly
before
our

48 publication.19‐21
 
All
our
ideas
were
based
on
the
general
fact
that
most
amino
acids

49 form
racemic
compound
crystals
that
are
less
soluble,
and
higher
melting,
than
the

50
crystals
of
the
L
or
the
D
enantiomers.

In
such
racemic
crystals
neighboring
L
and
D

51
52 molecules
interact
to
lower
the
free
energy—either
by
better
crystal
packing
or
by

53 direct
 stabilizing
 interactions
 in
 the
 crystals—compared
 with
 crystals
 where
 an
 L

54 molecule
has
only
other
L
neighbors.

55 




56
The
 idea
 is
 that
 a
 mixture
 of
 D
 and
 L
 amino
 acids
 with
 some
 excess
 of
 the
 L

57
58 component
(or
elsewhere
in
the
universe
perhaps
the
D
component)
would
dissolve

59
60

 5

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 6 of 15

1
2
3
4
in
 water,
 and
 as
 the
 water
 evaporates
 the
 less
 soluble
 racemate
 would
 precipitate

5 leaving
 a
 solution
 with
 increased
 richness
 in
 the
 L
 component.
 
 Because
 the

6 precipitation
 of
 the
 racemate
 depends
 on
 the
 solubility
 product
 SP(DL)
 =
 [D][L],

7 while
that
of
the
homochiral
L
compound
depends
only
on
the
solubility
[L],
there

8 would
 be
 feedback
 to
 increase
 the
 precipitation
 of
 the
 racemate
 as
 the

9
10
concentration
of
L
increases.

Thus
starting
with
a
small
excess
of
the
L
amino
acid,

11 the
 final
 L/D
 ratio
 would
 become
 very
 large
 with
 even
 a
 modest
 difference
 in

12 solubilities.

The
Morowitz
treatment17
is
shown
below.

13 

14 S(L)
=
[L]

15
16
SP(DL)
=
[D][L]

17 [D]
=
SP(DL)/[L]

18 [L]/[D]
=
S(L)2/SP(DL)

19 

20 The
 solubility
 product
 SP(DL)
 is
 equal
 to
 the
 square
 of
 half
 the
 solubility
 of
 the

21
22 racemate.

By
this
equation,
if
the
homochiral
crystals
were
only
twice
as
soluble
as

23 were
the
racemates
the
final
ratio
[L]/[D]
would
be
16/1,
a
solution
with
94%
of
the

24 L
 and
 6%
 of
 the
 D.
 If
 the
 homochiral
 crystals
 were
 three‐fold
 as
 soluble
 as
 the

25 racemate
the
final
L/D
ratio
would
be
36/1.
As
we
will
describe
later,
this
treatment

26 is
 not
 quite
 correct
 and
 can
 overestimate
 the
 selectivity
 for
 reasons
 we
 have

27
28 demonstrated.

29 


30 The
previous
theories
and
experiments
involved
solutions
at
equilibria,
but
kinetics

31 can
often
afford
different
results.

With
equilibrium
solubilities
we
saw
L
tryptophan

32
amplified
 to
 94.5%
 L
 and
 5.5%
 D,
 starting
 with
 any
 arbitrarily
 small
 excess
 (the

33
34 treatment
is
true
for
any
initial
ratio,
limited
only
by
the
practical
need
to
achieve

35 saturation
in
both
the
racemate
and
homochiral
compound,
so
smaller
amounts
of

36 water
would
be
needed
with
smaller
initial
L
excess).15

However,
we
found
that
this

37 ratio
 was
 raised
 to
 99/1
 in
 solution
 when
 we
 poured
 a
 small
 amount
 of
 water

38
through
 the
 final
 
 dry
 mixture,
 imitating
 the
 effect
 of
 rainwater.15
 Homochiral
 and

39
40 racemic
crystals
differ
in
their
activation
energies
for
subtracting
(or
adding)
units,

41 a
form
of
dissociation
(or
binding),
and
in
our
case
the
homochiral
crystals
dissolve

42 faster
beyond
the
requirements
of
the
equilibrium
constants.



43 

44
45
Forming
D
sugars
under
prebiotic
conditions
46 We
have
investigated
the
formation
of
the
simplest
sugar
belonging
to
the
D
series,

47 D‐glyceraldehyde.

All
other
sugars
in
the
D
family—including
D‐ribose,
D‐glucose,

48 D‐fructose,
 D‐erythrose,
 etc.—are
 named
 for
 the
 geometry
 of
 the
 chiral
 center

49 furthest
from
the
carbonyl
group
in
the
sugar,
and
this
is
the
unique
chiral
center
in

50
51
D‐glyceraldehyde.
 
 Sugars
 were
 probably
 synthesized
 on
 prebiotic
 Earth
 by
 a

52 process
 called
 the
 formose
 reaction.22
 D‐ribose
 (C5H1005)
 is
 a
 formal
 pentamer
 of

53 formaldehyde
 (CH2O),
 whose
 dimer
 is
 glycolaldehyde
 and
 whose
 trimer
 is

54 glyceraldehyde.
 
 In
 the
 formose
reaction
formaldehyde
 is
 treated
with
a
 mild
 base

55 such
as
calcium
hydroxide.

There
is
a
period
when
nothing
happens
until
there
is

56
57
the
 sudden
 conversion
 of
 the
 formaldehyde
 to
 glycolaldehyde
 and
 glyceraldehyde

58
59
60

 6

ACS Paragon Plus Environment
Page 7 of 15 Journal of the American Chemical Society

1
2
3
4
and
 even
 larger
 sugars.
 We
 proposed
 and
 demonstrated
 the
 mechanism
 of
 the

5 formose
reaction
(Figure
3)
by
an
autocatalytic
cycle
many
years
ago.23


6
7 slow HO
8 2 H 2C O H2C CH=O
9 H2C O
10
11
12 HO H
13 2 H2C CH=O HO CH=O
14
15 CH2OH
16
17
18 HOH
19
HOHC CH=O
20
CH2OH O
21
22 HOH2C CH2OH
23
24
25 O
H2C O
26 HOHC CH2OH
27
CH2OH 

28
29 

30 Figure
3.

The
formose
reaction,
producing
sugars
from
formaldehyde.

31 

32
33
A
first
molecule
of
glycolaldehyde
is
formed
from
formaldehyde
by
a
slow
unknown

34 process,
 possibly
 involving
 ionization
 by
 cosmic
 rays,
 and
 this
 then
 reacts
 with
 an

35 additional
formaldehyde
to
form
glyceraldehyde.


This
is
then
in
equilibrium
with

36 dihydroxyacetone
by
enolization
and
ketonization,
and
this
then
reacts
with
another

37 formaldehyde
 to
 form
 a
 four‐carbon
 ketosugar.
 
 By
 enolization
 and
 aldehyde

38
39
formation
 a
 four‐carbon
 aldehyde
 is
 formed
 that
 can
 undergo
 a
 reverse
 aldol

40 reaction
to
form
two
molecules
of
glycolaldehyde.


After
this,
four
glycolaldehydes

41 result
from
another
turn
of
the
cycle,
etc.



42 

43 We
 examined
 the
 formation
 of
 glyceraldehyde,
 a
 three‐carbon
 sugar
 that
 is
 the

44
45
simplest
 one
 with
 a
 chiral
 center.24
 Glyceraldehyde
 is
 formed
 in
 the
 formose
 cycle

46 by
the
reaction
of
formaldehyde
with
glycolaldehyde.


Without
a
chiral
catalyst

this

47 reaction
would
form
both
D‐glyceraldehyde
and
its
enantiomer
L‐glyceraldehyde
in

48 equal
amounts,
but
we
studied
what
would
happen
if
the
reaction
were
catalyzed
by

49 a
chiral
amino
acid.

We
can
trace
the
formation
of
L
amino
acids
back
to
the
small

50
51 excesses
of
S
α‐methyl
amino
acids
in
the
Murchison
meteorite,
so
the
question
was

52 simple.
 
 Would
 L
 amino
 acids
 catalyze
 the
 formation
 of
 D‐glyceraldehyde

53 preferentially,
at
least
to
a
small
extent,
and
if
so
could
the
preference
be
amplified

54 to
 a
 sufficient
 excess
 that
 the
 dominant
 D‐glyceraldehyde
 would
 be
 selected
 by

55
56
incipient
 life?
 
 If
 so,
 all
 of
 the
 D
 sugars—built
 on
 the
 D‐glyceraldehyde
 by
 adding

57 pieces
to
its
aldehyde
group
that
do
not
change
the
geometry
of
the
D
center
in
the

58
59
60

 7

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 8 of 15

1
2
3
4
glyceraldehyde
original
unit—are
part
of
the
general
origin
of
homochirality
tracing

5 back
to
the
meteorites.

If
not,
a
new
source
of
homochirality
will
need
to
be
found

6 for
the
sugars.

As
Table
1
shows,
we
examined
various
L‐amino
acids
from
different

7 classes.

8 

9
10

11 Table
1.

Ratio
of
D
to
L
glyceraldehyde

12 
from
glycolaldehyde
and
formaldehyde


13 catalyzed
by
various
L‐amino
acids.



14 The
reaction
conditions
and
analytical


15
16
method
are
described
in
ref.
24.

17 _________________________________

18 Serine
(50.3/49.7)

Alanine
(50.8/49.2)

19 Phenylalanine
(52.2/47.8)
Valine
(52.2/47.8)

20 Glutamic
acid
(60.7/39.3)
Proline
(28.9/71.1)
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

 8

ACS Paragon Plus Environment
Page 9 of 15 Journal of the American Chemical Society

1
2
3
4

5 

6 The
 reaction
 of
 formaldehyde
 with
 glycolaldehyde
 is
 an
 aldol
 reaction,
 and
 many

7 such
 reactions
 are
 known
 to
 be
 catalyzed
 by
 amines.
 
 They
 would
 react
 with
 the

8 glycolaldehyde
to
form
an
enamine,
and
this
adds
to
the
other
carbonyl
component.


9
10
Thus
 we
 carried
 out
 the
 reaction
 of
 glycolaldehyde
 with
 formaldehyde
 in
 water
 in

11 the
presence
of
a
variety
of
L
amino
acids.

We
found
(Table
1)
that
all
the
L
amino

12 acids
 caused
 the
 formation
 of
 glycolaldehyde
 with
 a
 small
 excess
 of
 the
 D

13 enantiomer,
with
one
exception:

L‐proline
catalyzed
the
formation
of
an
excess
of
L‐
14 glyceraldehyde.

(Some
recent
work
by
Blackmond25
indicates
that
in
basic
solution

15
16
this
proline
preference
is
reversed
when
the
carboxyl
group
is
a
carboxylate
ion,
so

17 proline
may
not
be
an
exception
after
all.)

18 

19 The
 preferences
 were
 modest.
 There
 was
 a
 60/40
 ratio
 of
 D/L
 glyceraldehyde

20 catalyzed
by
L‐glutamic
acid,
for
instance,
and
smaller
ratios
with
the
other
L‐amino

21
22 acids.
 
 Proline
 was
 probably
 not
 present
 in
 large
 amounts
 among
 the
 early
 amino

23 acids,
since
it
is
formed
in
two
secondary
reactions
from
glutamic
acid,
so
it
would

24 not
 dominate
 the
 chiral
 selectivity.
 
 The
 results
 with
 the
 other
 amino
 acids
 did

25 support
 the
 idea
 that
 the
 preference
 for
 D‐glyceraldehyde—and
 then
 the
 other
 D

26 sugars
derived
from
it—was
simply
the
result
of
the
formation
of
the
other
L
amino

27
28 acids
on
prebiotic
Earth.

29 



30 We
then
asked
whether
there
was
a
likely
process
for
amplifying
the
small
excess
of

31 D‐glyceraldehyde
 formed
 in
 this
 way,
 preferably
 by
 using
 the
 selective
 solubilities

32
that
 had
 worked
 with
 the
 amino
 acids.
 
 This
 turned
 out
 to
 be
 possible..
 
 D‐
33
34 Glyceraldehyde
 is
 a
 syrup
 with
 essentially
 complete
 water
 solubility,
 but
 racemic

35 DL‐glyceraldehde
 is
 a
 solid
 with
 a
 melting
 point
 of
 145°C
 and
 a
 limited
 water

36 solubility.

The
striking
difference
has
an
explanation.

A
previously
published
X‐ray

37 structure
determination
showed
that
DL‐glyceraldehyde
exists
as
a
chair‐form
six‐
38
membered
ring
dioxane
dimer
of
one
D
and
one
L
molecule,
with
all
the
substituents

39
40 equatorial
 as
 in
 chair
 cyclohexane,
 so
 this
 flat
 dimer
 molecule
 packs
 well
 into
 a

41 crystal.26
 
 With
 the
 same
 structure
 based
 on
 a
 dimer
 with
 D‐glyceraldehyde
 alone,

42 one
large
group
in
the
six‐membered
DD
ring
would
be
axial,
making
the
dimer
less

43 stable
and
also
making
crystal
packing
less
favorable.

The
result
was
that
we
could

44
45
take
the
60/40
ratio
of
D/L
glyceraldehyde
formed
by
catalysis
with
glutamic
acid

46 and
 turn
 it
 into
 a
 92/8
 D/L
 ratio
 in
 water
 solution
 by
 slow
 evaporation
 of
 water,

47 causing
the
racemic
crystals
to
precipitate.

48 


49 The
formation
and
amplification
of
D­ribonucleosides.



50
51
An
 aldol
 condensation
 of
 D‐glyceraldehyde
 wth
 glycolaldehyde
 would
 lead
 to
 D‐
52 ribose.
 
 We
 are
 studying
 the
 selective
 catalysis
 of
 this
 reaction,
 in
 which
 two
 new

53 chiral
centers
are
produced.

In
the
mean
time,
we
examined
the
possibility
that
any

54 excess
 of
 D
 over
 L
 in
 ribose
 could
 be
 amplified
 by
 selective
 solubilities,
 as
 in
 the

55 amino
acid
and
glyceraldehyde
cases.
In
contrast
to
the
situation
with
amino
acids,

56
57
we
 saw
 that
 D
 ribose
 and
 DL
 ribose
 had
 the
 same
 melting
 point.27
 
 Ribose

58 apparently
 forms
 a
 racemate
 that
 is
 a
 solid
 solution,
 with
 essentially
 identical

59
60

 9

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 10 of 15

1
2
3
4
properties,
solubilities,
and
melting
points
at
all
D/L
ratios.

In
the
solids
a
D‐ribose

5 molecule
 can
 equally
 well
 have
 either
 a
 D
 or
 an
 L
 as
 its
 neighbor.
 
 We
 did
 not

6 examine
 their
 solubility,
 since
 in
 all
 cases
 that
 we
 know
 of
 a
 lower
 solubility

7 correlated
 with
 a
 higher
 melting
 point—both
 melting
 and
 dissolving
 break
 up
 the

8 crystals.
 For
 this
 reason,
 D‐ribose
 cannot
 be
 amplified
 by
 the
 selective
 solubility

9
10
method.

Thus
we
examined
the
ribonucleosides
(Figure
5).27

11 


12 It
 had
 been
 shown
 previously
 that
 ribose
 reacts
 with
 purines
 under
 prebiotic

13 conditions
to
form
ribonucleosides.28
 
In
the
first
work
pyrophosphate
was
present

14 to
activate
the
ribose,
but
in
later
work
it
was
shown
that
the
pyrophosphate
could

15
16
be
replaced
by
the
residue
from
dried
seawater.

We
are
working
on
extending
these

17 results
 to
 pyrimidine
 nucleosides
 (see
 another
 proposal29
 on
 how
 pyrimidine

18 nucleosides
 could
 have
 been
 formed),
 and
 in
 the
 meantime
 we
 examined
 whether

19 ribonucleosides
 with
 small
 excesses
 of
 the
 D
 form
 could
 be
 amplified
 by
 the

20 solubility
method.

21
22 


23 NH2 O
24 N NH
N
25
26 N N N O
27 HO HO
28 O O
29 H H H H
30 H H H H
OH OH OH OH
31
32
Adenosine Uracil
33
34
35
36 NH2 O
37
N N NH
38
39
N O N N NH2
40 HO
HO
41 O
O
42 H H H H
43 H H H H
OH OH OH OH
44
45
46 Cytidine Guanosine
47
48
49

50 

51 Figure
4.

The
nucleosides
that
were
examined.

52 

53 We
 synthesized
 L‐uridine
 by
 a
 literature
 method
 and
 made
 a
 1:1
 mixture
 with
 D‐
54
55
uridine.
 We
 then
 crystallized
 the
 racemate
 to
 purity
 and
 examined
 its
 properties.


56 The
crystals
of
uridine
racemate
had
a
melting
point
of
176°C,
higher
by
21°C
than

57 the
 155°C
 for
 D‐uridine.
 
 D‐Uridine
 had
 a
 water
 solubility
 at
 22°C
 of
 454
 mg/mL,

58
59
60

 10

ACS Paragon Plus Environment
Page 11 of 15 Journal of the American Chemical Society

1
2
3
4
while
the
racemic
uridine
had
a
solubility
of
only
87
mg/mL,
5.2
times
smaller.

By

5 the
Morowitz
calculation
this
should
afford
a
D/L
ratio
of
108/1
at
saturation
with

6 both
crystals.

We
then
dissolved
both
the
D‐uridine
and
the
DL‐uridine
crystals
to

7 saturation
together
in
water
at
22°C,
filtering
away
the
excess,
and
saw
96%
D
and

8 4%
L
in
solution,
a
ratio
of
24/1.

The
large
ratio
makes
the
D‐uridine
dominant,
but

9
10
it
is
less
than
predicted
by
the
Morowitz
theoretical
treatment.

11 

12 We
 then
 examined
 the
 situation
 with
 adenosine,
 synthesizing
 L‐adenosine
 and

13 mixing
 it
 with
 D‐adenosine,
 then
 recrystallizing
 the
 racemic
 crystals
 to
 purity.
 
 In

14 this
case
the
melting
point
of
the
racemate
was
243
±
1°C,
while
for
D‐adenosine
the

15
16
m.p.
was
230
±
1°C.

The
solubility
of
D‐adenosine
in
water
at
22°C
was
5.2
mg/mL,

17 6.3
 times
 as
 large
 as
 the
 0.8
 mg/mL
 for
 the
 DL
 crystal.
 There
 should
 have
 been
 a

18 160/1
 D/L
 ratio
 at
 saturation
 equilibrium
 from
 the
 Morowitz
 treatment,
 but
 we

19 observed
 a
 99/1
 ratio
 of
 D/L
 adenosine.
 
 Again
 very
 large,
 but
 again
 less
 than

20 predicted
 by
 the
 Morowitz
 treatment.
 
 
 With
 cytidine
 we
 saw
 decomposition
 on

21
22 melting
of
both
the
D
and
the
DL
crystals,
so
no
melting
points
could
be
determined,

23 but
 the
 solubilities
 predicted
 an
 even
 larger
 amplification
 than
 for
 the
 other
 two

24 nucleosides.
 D‐cytidine
 had
 a
 solubility
 of
 192
 mg/mL
 in
 22°C
 water,
 while
 DL‐
25 cytidine
 had
 a
 solubility
 of
 24.3
 mg/mL.
 
 The
 Morowitz
 equation
 predicts
 D/L
 of

26 250/1
at
saturation,
while
we
observed
199/1.


27
28 


29 All
three
of
these
nucleosides
can
be
amplified
to
very
high
D/L
ratios
by
selective

30 solubilities,
 but
 the
 Morowitz
 treatment
 overestimated
 the
 ratios.
 
 We
 concluded

31 that
 this
 could
 reflect
 the
 fact
 that
 the
 solubilities
 being
 measured
 were
 in
 pure

32
water,
 but
 the
 solubilities
 relevant
 to
 the
 amplification
 experiments
 are
 in
 water

33
34 along
 with
 a
 second
 component.
 
 In
 the
 experiment,
 the
 solubility
 of
 the
 racemate

35 that
is
relevant
is
that
in
the
presence
of
the
homochiral
component
that
is
also
in

36 solution.
 
 We
 concluded
 that
 the
 homochiral
 dissolved
 material
 was
 acting
 as
 a

37 cosolvent,
 increasing
 the
 solubility
 of
 the
 racemate
 over
 that
 in
 pure
 water.14
 
 The

38
dissolved
homochiral
component
has
two
roles.

In
one
it
helps
drive
the
racemate

39
40 out
of
solution
by
its
role
in
the
solubility
product
of
the
racemate,
but
in
the
other

41 role
 it
 increases
 the
 solubility
 of
 the
 racemate
 by
 acting
 as
 an
 antihydrophobic

42 cosolvent
 like
 ethanol,
 with
 which
 relatively
 hydrophobic
 substrates
 such
 as

43 nucleosides
have
greater
solubility
than
in
pure
water.

If
this
is
true
the
solubility
of

44
45
the
homochiral
crystals
would
also
be
increased
by
the
presence
of
the
racemate
in

46 solution,
 but
 at
 final
 saturation—with
 little
 dissolved
 racemate—this
 effect
 would

47 be
smaller.


48 



49 To
 test
 this
 idea,
 we
 examined
 the
 solubility
 of
 racemic
 uridine
 in
 water
 with
 D‐
50
51
cytidine
added
to
saturation.14

This
can
mimic
the
cosolvent
effect
of
D‐uridine,
but

52 does
 not
 play
 a
 role
 in
 the
 solubility
 product
 of
 DL‐uridine.
 
 We
 found
 that
 the
 D‐
53 cytidine
increased
the
solubility
of
DL‐uridine
by
40%
over
that
in
pure
water.
Thus

54 the
Morowitz
treatment
overestimates
the
amplifications
a
bit,
because
it
does
not

55 include
 the
 cosolvent
 effect.
 
 The
 experimental
 values
 are
 still
 so
 high
 that
 they

56
57
could
lead
to
the
dominance
of
the
D‐nucleosides
on
prebiotic
Earth,
even
starting

58 from
a
very
small
initial
excess.

59
60

 11

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 12 of 15

1
2
3
4

5 Guanosine
 behaved
 differently.27
 
 D‐guanosine
 and
 the
 DL‐guanosine
 crystals

6 melted
 with
 decomposition,
 but
 we
 could
 still
 examine
 their
 solubilities
 in
 22°C

7 water.
 
 DL‐Guanosine
 had
 a
 solubility
 of
 0.84
 mg/mL
 while
 D‐guanosine
 had
 a

8 solubility
of
0.46
mg/mL,
essentially
half
that
of
the
racemate.


This
indicated
that

9
10
the
racemate
belongs
to
the
third
class
of
racemic
crystals.

It
is
not
a
solid
solution

11 as
 we
 saw
 with
 ribose,
 nor
 a
 racemic
 compound
 crystal
 like
 those
 of
 uridine,

12 adenosine
and
cytidine.

It
is
a
racemic
conglomerate
of
D
and
L
crystals,
each
with

13 its
own
solubility.

Such
a
conglomerate
is
like
that
with
racemic
sodium
ammonium

14 tartrate,
that
allowed
Pasteur
to
separate
the
D
and
L
crystals
by
hand.

15
16




17 This
 inversion
 in
solubilities,
 with
 the
 racemate
more
soluble
 than
 the
homochiral

18 compound,
makes
it
impossible
to
amplify
D‐guanosine
by
selective
solubilities,
but

19 of
 course
 the
 D‐ribose
 from
 hydrolysis
 of
 the
 other
 three
 nucleosides
 could
 have

20 been
 used
 to
 synthesize
 D‐guanosine
 on
 prebiotic
 Earth.
 
 The
 D‐ribose
 resulting

21
22 from
hydrolysis
of
the
other
three
nucleosides
could
also
play
additional
important

23 biochemical
roles,
including
conversion
to
D‐ribulose
whose
diphosphate
is
the
key

24 sugar
in
photosynthesis.
25
26
27 
Conclusion.

This
work30,
31
answers
some
of
the
questions
in
the
general
idea
that

28 the
unusual
amino
acids
delivered
to
Earth
by
the
Murchison
meteorite
and
related

29 ones
could
have
led
to
the
dominance
of
L
amino
acids
and
D
sugars
on
early
Earth

30 that
would
permit
life
to
start.

Of
course
showing
that
it
could
have
happened
this

31
32 way
is
not
the
same
as
showing
that
it
did.

Proper
theories
need
the
possibility
of

33 falsification.
This
account
would
be
in
trouble
if
significant
examples
were
found
in

34 which
meteorites
that
landed
on
Earth
contained
R
α‐methyl
amino
acids
instead
of

35 the
 S
 examples
 in
 the
 Murchison
 meteorite
 and
 others
 examined
 so
 far.
 
 
 An

36
alternative
 scenario
 to
 ours
 would
 use
 the
 Murchison
 α‐methyl
 amino
 acids
 to

37
38 catalyze
the
formation
of
some
D
sugars,
as
Pizzarello
and
Weber
have
shown32
(cf.

39 refs.
 (33)
 and
 (34)
 for
 related
 work),
 and
 then
 use
 those
 sugars
 to
 catalyze
 the

40 formation
 of
 the
 normal
 proteinogenic
 L
 amino
 acids.
 
 Good
 credibly‐prebiotic

41 examples
of
the
latter
process
have
not
yet
been
produced.


42
43
44 Further
 work
 is
 needed
 to
 show
 prebiotic
 versions
 of
 the
 conversion
 of
 D‐
45 glyceraldehyde
 to
 D‐ribose,
 D‐glucose,
 D‐fructose,
 and
 D‐2‐deoxyribose,
 efforts

46 underway
 in
 our
 lab.
 
 We
 also
 need
 a
 credible
 way
 in
 which
 nucleosides
 and

47 nucleotides
 could
 be
 formed
 from
 these
 sugars
 and
 pyrimidines.,
 not
 just
 purines


48 Finally,
 of
 course,
 we
 all
 need
 ways
 in
 which
 these
 and
 other
 sensible
 building

49
50 blocks
could
assemble
into
structures
with
the
exciting
properties
of
life.

51 


52 
An
implication
from
this
work
is
that
elsewhere
in
the
universe
there
could
be
life

53 forms
 based
 on
 D
 amino
 acids
 and
 L
 sugars,
 depending
 on
 the
 chirality
 of
 circular

54 polarized
light
in
that
sector
of
the
universe
or
whatever
other
process
operated
to

55
56 favor
the
L
α‐methyl
amino
acids
in
the
meteorites
that
have
landed
on
Earth.

Such

57 life
 forms
 could
 well
 be
 advanced
 versions
 of
 dinosaurs,
 if
 mammals
 did
 not
 have

58
59
60

 12

ACS Paragon Plus Environment
Page 13 of 15 Journal of the American Chemical Society

1
2
3
4
the
 good
 fortune
 to
 have
 the
 dinosaurs
 wiped
 out
 by
 an
 asteroidal
 collision,
 as
 on

5 Earth.

We
would
be
better
off
not
meeting
them.

6
7
Acknowledgements

8
9
10 I
thank
my
coworkers
whose
names
appear
in
the
references,
and
the
U.
S.
National

11 Science
Foundation
for
support
of
this
work.


12
13 

14 

15
16

17 References

18 1.
Pizzarello,
S.;
Groy,
T.

Geochim.
Cosmochim.
2011,
75,
645‐656.

19 2.
Kvenvolden,
K.;
Lawless,
J.;
Pering,
K.;
Peterson,
E.;
Flores,
J.;
Ponnamperuma,
C;

20 Kaplan,
I.;
Moore,
C.

Nature
1970,
228,
923‐926.


21
22
3.
Cronin,
J.;
Pizzarello,
S.

Science
1997,
275,
951‐955.

23 4.
Degenkolb,
T.;
Gams,
W.;
Bruckner,
H.

Chem.
&
Biodiversity
2008,
5,
693‐706.


24 5.
Degenkolb,
T.;
Bruckner,
H.

Chem.
&
Biodiversity
2008,
5,
1817‐43.


25 6.
Bruckner,
H.;
Becker,
D.;
Gams,
W.;
Degenkolb,
T.

Chem.
&
Biodiversity
2009,
6,

26 38‐56.

27
28
7.


Herbst,
E.;
van
Dishoek,
E.
Ann
Rev.
Phys.
Astron.
&
Chem.
2009,
47,
427‐480.

29 8.
Strecker,
A.

Annal.
Chem.
Pharm.
1850,
75,
27‐45.

30 9.
Glavin,
D.;
Dworkin,
J.

Proc.
Natl.
Acad.
Sci.
USA
2009,
106,
5487–5492.

31 10.
Flores,
J.;
Bonner,
W.;
Massey,
G.

J.
Am.
Chem.
Soc.
1977,
99,
3622‐3625.

32 11.
Bailey,
J.;
Chrysostomou,
A.;
Hough,
J.;
Gledhill,
T.;
McCall,
A.;
Clark,
S.;
Menard,
F.;

33
34 Tamura,
M.

Science
1998,
281,
672‐674.

35 12.
Bailey,
J.

Orig.
Life
Evol.
Biosph.
2001,
31,
167‐183.

36 13.
Buschermohle,
M.;
Whittet,
D.;
Chrysostomou,
A.;
Hough,
J.;
Adamson,
A.;

37 Whitney,
B.;
Wolff,
M.

Astrophys.
J.
2005,
624,
821‐826.

38 14.
Breslow,
R.;
Levine,
M.;
Cheng,
Z.

Orig.
Life
Evol.
Biosph.
2010,
40,
11‐26.

39
40 15.
Levine,
M.;
Kenesky,
C.;
Mazori,
D.;
Breslow,
R.

Org.
Lett.
2008,
10,
2433‐2436.

41 16.
Chruma,
J.;
Liu,
L.;
Zhou,
W.;
Breslow,
R.

Bioorg.
Med.
Chem.
2005,
13,
5873‐
42 5883.


43 17.
Morowitz,
H.

J.
Theor.
Biol.
1969,
25,
491.

44
18.
Breslow,
R.;
Levine,
M.

Proc.
Natl.
Acad.
Sci.
USA
2006,
103,
12979‐12980.

45
46 19.
Klussmann,
M.;
Iwamura,
H.;
Mathew,
S.;
Wells,
D.;
Pandya,
U.;
Armstrong,
A.;

47 Blackmond,
D.

Nature
2006,
441,
621.

48 20.
Klussmann,
M.;
Izumi,
T.;
White,
A.;
Armstrong,
A.;
Blackmond,
D.

J.
Am.
Chem.

49 Soc.
2007,
129,
7657‐7660.

50
21.
Noorduin,
W.;
Izumi,
T.;
Millemaggi,
A.;
Leeman,
M.;
Meekes,
H.;
Van
Enckevort,

51
52 W.;
Kellogg,
R.;
Kaptein,
B.;
Vlieg,
E.;
Blackmond,
D.

J.
Am.
Chem.
Soc.
2008,
130,

53 1158‐1159.

54 22.
Langenbeck,
W.

Tetrahedron
1958,
3,
185‐196.

55 23.
Breslow,
R.

Tetrahedron
Lett.
1959,
1
(21),
22‐26.

56
57
24.
Breslow,
R.;
Cheng,
Z.

Proc.
Natl.
Acad.
Sci.
USA
2010,
107,
5723‐5725.



58 25.
Hein,
J.;
Blackmond,
D.
G.
Accts.
Chem.
Res.
2012,
30,
xxxx.


59
60

 13

ACS Paragon Plus Environment
Journal of the American Chemical Society Page 14 of 15

1
2
3
4
26.
Senma,
M.;
Taira,
Z.;
Osaki,
K.;
Taga,
T.

J.
Chem.
Soc.
Chem.
Commun.
1973,
880‐
5 881.

6 27.
Breslow,
R.;
Cheng,
Z.

Proc.
Natl.
Acad.
Sci.
USA
2009,
106,
9144‐9146.

7 28.
Fuller,
W.
D.;
Sanchez,
R.
A.;
Orgel,
L.
E.
J.
Mol.
Evoln.
1972,
1,
249‐257.
W. D. 

8 29.
Powner,
M.
W.;
Gerland,
B.;
Sutherland,
J.
D.
Nature
2009,
459,
239–242.

9
10
30.
Breslow,
R.

Tetrahedron
Lett.
2011,
52,
2028‐2032.

11 31.
Breslow,
R.

Israel
J.
Chem.,
2011,
51,
1‐7.

12 32.
Pizzarello,
S.;
Weber,
A.

Orig.
Life
Evol.
Biosph.
2010,
40,
3‐10.

13 33.
Burroughs,
L.;
Vale,
M.;
Gilks,
J.;
Hayes,
H.;
Christopher,
J.;
Clarke,
P.

Chem.Comm.

14 2010,
46,
4776‐4778.

15
16
34.
Fernandez‐Lopez,
R.;
Kofoed,
J.;
Machuqueiro,
M.;
Darbre,
T.

Eur.
J.
Org.
Chem.

17 2005,
24,
5268‐5276.

18 

19 Figure
captions

20 

21
22 Figure
 1.
 
 Some
 α‐methyl
 amino
 acids
 discovered
 in
 the
 Murchison
 and
 other

23 meteorites,
 all
 of
 which
 have
 a
 small
 excess
 of
 the
 S
 configuration
 that
 was

24 described
 as
 L.
 
 The
 enantioexcesses
 showed
 a
 range
 of
 values
 in
 this
 and
 later

25 work.

26
27

28 Figure
2.

Transaminative
decarboxylation
of
the
imine
from
S‐α‐methylvaline
and

29 an
α
ketoacid.

In
the
presence
of
Cu(II),
which
catalyzes
the
reaction,
the
product

30 amino
acid
has
an
excess
of
the
L
enantiomer,
but
not
without
the
Cu(II).

This
is

31
32
expected
if
a
square
planar
complex
is
formed
between
two
imines
and
the
Cu(II),
so


33 after
decarboxylation
of
one
ligand
it
is
protonated
selectively,
guided
by
the

34 stereochemistry
of
the
other
ligand.


35 

36 Figure
3.

The
formose
reaction,
producing
sugars
from
formaldehyde.

37
38

39 

40 Figure
4.

The
nucleosides
that
were
examined
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

 14

ACS Paragon Plus Environment
Page 15 of 15 Journal of the American Chemical Society

1
2
3
4
Graphic
Abstract

5 

6 

7 

8 

9
10

11 

12 

13 

14 

15
16

17 

18 

19 

20 

21
22 

23 

24 

25 

26 

27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

 15

ACS Paragon Plus Environment

You might also like