45 IChO 2013 Prep Problems Solutions+Index
45 IChO 2013 Prep Problems Solutions+Index
2) 25% of carbon atoms retain the sp2 hybridization, which means that they are not bonded
to oxygen atoms. 75% of carbon atoms form chemical bonds with oxygen. Each oxygen atom is
bonded to the pair of carbon atoms. The net formula is СО0.375. Maximum Х in the Hoffman
model is 0.5. The net formula is СО0.5.
3) The four groups are the phenol (OH sp2), hydroxyl (OH sp3), and epoxide groups in the
basal plane, and the carboxylic acid groups at the edges.
4) Each hydrogen atom corresponds to one oxidized carbon atom. 22% of carbon atoms are
bonded to the hydroxyl or phenol group, or are in the carboxylic acid group. Let all the hydrogen
atoms be in the carboxylic acid groups. Then 44% of oxygen atoms are in the carboxylic acid
groups and 2% are in the epoxy groups. In this case 22% + 2⋅2% = 26% of all the carbon atoms
are oxidized. 74% of the total amount of carbon atoms do not form chemical bonds with oxygen.
This is the upper limit. Let all the hydrogen atoms be in the hydroxyl or phenol groups. This
means that there are no carboxylic acid groups in the particular GO sample! Then 24% of
oxygen atoms are in the epoxy groups. In this case 22% + 2⋅24% =70% of all the carbon atoms
are bonded to oxygen. 30% of carbon atoms are not oxidized. This is the lower limit.
5) Acid groups do not participate in the hydrogen bonding network (Fig. 3). It means that
maximum degree of water absorption will be reached in case of the absence of such groups in
GO. Then each pair of hydrogen atoms holds one molecule of Н2О (0.11), and each pair of
epoxy groups also holds one molecule of Н2О (0.46–0.22) / 2 = 0.12. Altogether there are 0.23
molecules of water per one carbon atom. The chemical formula of GO hydrate is
СН0.22О0.46⋅0.23Н2О.
2
Problem 2. Efficiency of photosynthesis
The correction from non-standard pressures is not large – about 1/30 of the standard Gibbs
energy.
3. Energy of 10 mol of photons absorbed by green plants is 176⋅10 = 1760 kJ. Of this
amount 480.5 kJ is converted to Gibbs energy. The efficiency of the solar energy conversion by
green plants can be estimated as 480.5 / 1760 ⋅ 100% = 27%.
3. H2O2. The compound [Cr(NH3)5Cl]Cl2 is formed because the oxidation takes place via
the η2-bridging peroxocomplex, followed by the hydrolysis when the leaving peroxo-group is
replaced by the chloride-ion from the solution.
5. The chromium(3+) complexes are inert, thus the substitution process occurs slowly. This
is due to the d3 configuration.
See J.D. Atwood, Inorganic and organometallic reaction mechanisms, 2nd edition, Wiley-VCH,
pp.85-86 and P.C. Ford et al, Inorg. Chem., 1968, 7, 1976.
1. The common mineral of tin is cassiterite, SnO2. Thus, 1.05 g of Х after decomposition
give 0.8664 g of SnO2 that contains 5.738 mmol of tin. Under decomposition 0.069 g (3.833
mmol) of water form. As the ratio n(Sn) : n(H2O) is equal to 1.5 (or 3 : 2), the brutto formula of
Х contains 3 equivalents of SnO2, 4 of H and 2 of O (from 2 water molecules). In addition, it
also contains nitrogen and probably more oxygen. Their mass is 1.05 – 0.8664 – 0.069 = 0.1146
g and the average molar mass is M = 0.1146 / (0.00383/2) = 60 g/mol, which corresponds to
N2O2. Thus, the formula of X is Sn3O10N2H4, or Sn3O2(NO3)2(H2O)2.
2. All the operations should be performed in an inert atmosphere, because tin(II) hydroxide
is oxidized in air.
3. If all the metal atoms in the cation are equivalent they have the same coordination sphere.
So, we may suppose the formula [Sn3(OH)4]2+, that is a combination of three pyramids linked by
joint edges in a cycle (See J.D. Donaldson et al, JCS Dalton Trans, 1995, 2273.):
Sn
OH
HO OH
Sn
OH
Sn
5
The pyramidal nonplanar geometry is due to the electron pair on each tin atom.
4. In the acidic solution the hydrated tin(2+) ions are formed, in the basic media – the anions
[Sn(OH)3]–, [Sn(OH)6]4– and oligonuclear species such as [Sn2O(OH)4]2–, [Sn4O(OH)10]4–.
RT 8.314 ⋅ 298
2.
3. Y – [NS(O)F]3
2[NS(O)F]3 + 9Ba(CH3COO)2 + 18H2O = 3BaF2↓ + 6BaSO4↓ + 12CH3COOH + 6CH3COONH4
4. [NS(O)(NHCH3)]3
5. and (SO3)3
6. Z = (SO3)n
6
n(AB) 2 1.52
K1 = = = 7.2
n(A 2 )n(B2 ) 1.25 ⋅ 0.25
n(AB) 2 (2 y ) 2
K1 = =
n(A 2 )n(B2 ) (1 − y ) ⋅ (1 − y )
n(AB) 2y y K1
= = = = 1.34
n(A 2 ) + n(B2 ) (1 − y ) + (1 − y ) 1 − y 4
5. a) At x→∞ the very large amount of A2 will almost completely shift the equilibrium
A2 + B2 = 2AB to the right, and almost all B2 will be converted to AB, the yield will tend to 1,
η( x → ∞) → 1 .
b) At x→ 0 (1/x→∞) the situation is the same as in (a) if we interchange A2 and B2, that
is η( x → 0) → 1 .
6. From question 5 it follows that at x = 1 the function η(x) has a minimum, because at x = 0
or x = ∞ it approaches the maximum possible value of 1. Qualitatively, the graph is as follows:
η(x)
0
0 1
x
7. Suppose we have in total 1 mol of A2 and B2, and the molar ratio A2 : B2 = x : 1. Then,
the initial amounts of reagents are: n(A2) = x/(x+1), n(B2) = 1/(x+1). It follows from the
symmetry between A2 and B2 that the equilibrium amount of AB will be the same for the molar
ratios x and 1/x, hence x = 1 corresponds to the maximum or minimum neq(AB).
If x is very large (small), then the initial amount of B2 (A2) will be small and so will be
neq(AB). Therefore, the maximum amount of AB will be obtained at A2 : B2 = 1 : 1. The
equilibrium calculation for this case is as follows.
9
0.5 0.5
A2 + B2 = 2AB
y y 2y
0.5–y 0.5–y 2y
n(AB) 2 (2 y ) 2
K= =
n(A 2 )n(B2 ) (0.5 − y ) 2
K
y=
4+2 K
K
neq (AB) =
2+ K
1. CuSO4·5H2O.
∆H d 1 1
=ph ph 0 exp − ,
2R T0 T
where ph0 = 1047 Pa is the saturated vapor pressure over CuSO4·5H2O and T0 = 298К. Enthalpy
of decomposition of CuSO4·5H2O is: ∆Hd = 2·(–241.83) – 1688.7 + 2277.4 = 105.04 kJ·mol–1.
The similar equation describes the temperature dependence of the vapor pressure of water
pw:
∆H vap 1 1
=pw pw 0 exp − .
R T0 T
The enthalpy of vaporization of water is: ∆Hvap = –241.83 + 285.83 = 44.0 kJ·mol–1. The
humidity is the ratio of two vapor pressures:
ph ph 0 ∆H d / 2 − ∆H vap 1 1
= exp = − 0.35 .
pw pw 0 R T0 T
10
From this equation we find the required temperature:
1 1 R 0.35 pw 0 1 8.314 0.35 ⋅ 3200
= − ln = − ln =
0.00329
T T0 ∆H d / 2 − ∆H vap ph 0 298 (105.04 / 2 − 44 ) ⋅ 10 3
1047
1
=T0 = 304 K or 31 °С.
0.00329
3. b)
4. After several repetitions of the procedure, the equilibrium is established between the
anhydrous copper sulfate and its monohydrate: CuSO4·H2O = CuSO4 + H2O. In this case the
saturated vapor pressure of water over its solution in ethanol is equal to the saturated vapor
ph
pressure of water over CuSO4·H2O. Thus, p= p w γx =
, x = 0.0136 , the mass fraction of
pw γ
h
water is:
xM (H 2O)
=w(H 2O) = 0.0054 , or 0.54%.
xM (H 2O) + (1 − x ) M (C2 H 5OH)
ph ph 0 ∆H d − ∆H vap 1 1
=x = exp − .
p w γ γp w 0 R T0 T
maximum efficiency of the catalyst is achieved. Now ∆NB/∆t is independent of the reagent
pressure and no more A can be converted into products per unit of time per catalytic site. The
initial slopes of the curves (10) and (11) should be used to calculate TOF and TON
∆N B
= 210 s −1 ; ТОN ≤ ТОF × t = 210 ⋅ 40 ⋅ 60 = 5 ⋅ 105
∆t ⋅ 1015
3. a) The slope of the linear dependence in Fig. 2a should be used to calculate TOF:
= 6s −1
TOF= tan α
It is assumed that every single atom of the metal forms a catalytic site and works independently.
TOF is independent of the amount of atoms deposited.
b) In this case a group of n atoms, rather than a single atom, forms a catalytic site. The
number of catalytic sites is
4) The authors of this study considered every single Au atom to be a catalytic site. One has
to calculate the number of Au atoms involved in the catalytic process in Fig. 3a and 3b. In the
case (b), all yellow spheres are taking part in the reaction. In the case (a), 1/3 of the yellow
spheres from the lower monolayer are involved together with all red spheres. 2/3 of the yellow
12
spheres are blocked by the red spheres from the top and do not participate in the catalytic
reaction.
Let NAu be the number of the yellow spheres in Fig. 3b. The number of the red spheres in
Fig. 3a is equal to 1/3 NAu. The total number of Au atoms involved in catalytic reaction in Fig. 3a
is 1/3 NAu(red) + 1/3 NAu (yellow). The rate of the reaction in case (а) is:
1
=
r2 4=
r1 r2 (red) + r1 ,
3
where r2(red) and 1/3r1 are partial rates for the red and yellow spheres, respectively.
1
4r1 − r1
TOF (a) 3= r 3(11 / 3r1 )
Finally, = : 1 = 11
TOF (b) 1 N r1
N Au Au
3
Below are given the mechanisms of these reactions, established by various experimental
methods. However, the limited data given in the text of a problem allow multiple mechanisms.
Therefore the only two criteria for the correct solutions are: 1) the consistency of the mechanism
with the rate law; 2) the chemical sense.
2. The catalytic effect of silver is due to formation of silver(II) ions and sulfate ion radicals upon
reaction of Ag+ with persulfate. The mechanism is:
Ag+ + S2O82 – → ⋅SO4– + SO42– + Ag2+ slow, rate-determining
k2[HCOO–][⋅SO4–] = k3[⋅CO2–][S2O82–]
Hence
[⋅CO2–] = (k1k2)1/2(k3k4)–1/2[HCOO–]1/2
[⋅SO4–] = (k1k3)1/2(k2k4)–1/2[S2O82–][HCOO–]–1/2
The rate of the reaction is equal to the rate of formate consumption:
r = r2 = k2[HCOO–][⋅SO4–] = (k1k2k3)1/2k4–1/2[HCOO–]1/2[S2O82–] = keff[HCOO–]1/2[S2O82–]
A more complex mechanism includes the formation of OH radicals and several chain
termination reactions. That’s why the given rate law is valid only for a limited range of reactant
concentrations.
4. The rate-determining step is the addition of azide ion to the solvent, carbon disulfide:
N
N
CS2 N3 -
S-
N
S
The oxidation of this ion by iodine is a series of fast reactions. The overall rate of the reaction
I2 + 2N3– = 3N2 + 2I–
is half of that of the azide-CS2 reaction:
r = ½ k[N3–][CS2].
Introducing the effective constant keff = ½ k[CS2] we get:
r = keff[N3–].
5. The reaction mechanism includes several steps. The first step is the reversible addition of
DABCO to ester:
O
N COOR2 OR2
+
N
N
N
The next two steps are the reversible additions of two molecules of aldehyde to the zwitter ion
formed in the previous step:
15
O O O
OR2 R1 OR2
N + R1CHO
N
N N
R1 O
O O O O
H
R1 OR2 R1 OR2
+ R1CHO
N N
N N
R1 O
O O R1 OH
H
O O N
R1 OR2
+
N R1 OR2 N
N
After that, the product rapidly eliminates one molecule of aldehyde. Applying quasi-equilibrium
conditions to the first three steps, we get:
r = kRDSK1K2K3[aldehyde]2[ester][DABCO] = keff[aldehyde]2[ester][DABCO]
It is worth mentioning that in protic solvents the rate-determining step is the solvent-assisted
proton transfer in DABCO-ester-aldehyde adduct, hence the reaction order is one with respect to
either aldehyde, or ester or base.
6. The first step of the reaction is the reversible addition of peroxyacid anion to the carboxylic
group of peroxyacid:
16
O
O
R O
RCOOO + RCOOOH O H
O
O R
O
O
R O
O H RCOOH + RCOO + O2
O
O R
where Ka is the acidity constant of peroxyacid. Substituting these concentrations to the rate law
we obtain:
keff K a H + k1 H +
= ( RCO3H )
r c= 2
c ( RCO3 H )
2
( ) ( )
2 2
Ka + H +
k2 + H +
Note that at given c(RCO3H) the reaction rate is maximum if [RCO3H] = [RCO3–] (and [H+] =
Ka).
Since this is a reactor with ideal stirring, the concentrations of substances in the output
flow are equal to the concentrations inside the reactor. In a stationary state, the concentrations
and quantities of substances in the reactor are constant. Consider the material balance with
respect to X, Y and P.
Stationary conditions are:
17
∆ν X , R ∆ν Y , R ∆ν P , R
= 0= 0= 0, (1)
∆t ∆t ∆t
where ΔνX,R, ΔνY,R, ΔνP,R are the changes of the quantities for the substances X, Y and P in the
reactor during time Δt. The quantity of the substance in the reactor may change due to input
flow, chemical reaction, and output flow:
∆ν X , R ∆ν X , R ∆ν X , R ∆ν X , R
= + + (2)
∆t ∆t input ∆t reaction ∆t output
The same is true for Y and P.
Input flow rates of the substances are
∆ν X , R ∆ν ∆ν P , R
=
f X=
cX , I Y , R f=
Y cY , I 0, (3)
∆t input ∆t input ∆t input
where fX and fY are the input volumetric flows of the solutions of X and Y, cX,I and cY,I –
concentrations of X and Y in the respective solutions.
Let the balanced reaction equation be
nX X + nY Y = nP P
where nX, nY and nP are the stoichiometric coefficients for the corresponding substances. Due to
a chemical reaction the quantities of the substances in the reactor change with the rates
∆ν X , R ∆ν Y , R ∆ν P , R
=
−nX rVR =
−nY rVR =
nP rVR , (4)
∆t reaction dt reaction ∆t reaction
where r – the reaction rate, VR – the reactor volume.
The output flows of the substances are:
∆ν X , R ∆ν Y , R ∆ν P , R
=
f=
O cX , R f=
O cY , R f O cP , R , (5)
∆t output ∆t output ∆t output
where fO is the volumetric output flow, cX,R, cY,R and cP,R – the concentrations of substances X, Y
and P in the reactor. Since the process is stationary and the reaction proceeds in the liquid phase,
the output volumetric flow equals the sum of input volumetric flows:
f=
O f X + fY (6)
Thus the material balance equations (2) considering expressions (1) and (3)-(6) are
∆ν X , R
= f X c X , I − nX rVR − c X , R ( f X + fY =
) 0
∆t
∆ν Y , R
= fY cY , I − nY rVR − cY , R ( f X + fY )= 0
∆t
∆ν P , R
= nP rVR − cP , R ( f X + fY )= 0
∆t
Hence
18
nX rVR =f X c X , I − c X , R ( f X + fY )
nY rVR =fY cY , I − cY , R ( f X + fY )
nP rVR cP , R ( f X + fY )
=
One of the possible mechanisms that match the obtained rate law is:
X+Y↔I fast
I+Y→P slow, rate-determining
1. Since all reaction pathways obey the same rate law, the quantity of the product is
proportional to the respective rate constant. The overall constant equals the sum of constants for
different pathways. Hence
for 1,2-dichloro:
61%
k=
1.45 ⋅ 104 8.8 ⋅ 103 M −1s −1
=
100%
for 1-acetoxy-2-dichloro:
30%
k=
1.45 ⋅ 104 4.4 ⋅ 103 M −1s −1
=
100%
for 2-chlorostyrene:
9%
k=
1.45 ⋅ 104 1.3 ⋅ 103 M −1s −1
=
100%
2. This reaction is not stereospecific and leads to the formation of diastereomeric addition
products in comparable amounts. The following products are obtained (approximate ratios of
product quantities at 25oC are given as an illustration):
1,2-dichloro:
20
Cl Cl Cl Cl
Cl Cl Cl Cl
1-acetoxy-2-chloro:
OAc OAc OAc OAc
Cl Cl Cl Cl
2-chlorostyrene:
Cl
Cl
~1:3
The achiral sorbent is unable to separate enantiomers, so only 6 different fractions can be
obtained. The chiral sorbent allows full separation, so in this case the determined number of
products is 10.
1. The boiling point of water and the melting point of ice V increase, and the melting point
of ordinary ice decreases with the increasing pressure. This can be easily explained using the Le
Chatelier principle. In the phase transitions
H2O(l) H2O(g)
and
H2O(ice,V) H2O(l)
the volume increases and heat is absorbed (∆V > 0, ∆H > 0). Hence, with the increasing pressure
both equilibria are shifted to the left; consequently, temperature should be increased to keep the
equilibria.
In the phase transition
21
H2O(ice,I) H2O(l)
the volume decreases and heat is absorbed (∆V < 0, ∆H > 0). Hence, with the increasing pressure
the phase equilibrium is shifted to the right, and temperature should be decreased to keep the
equilibrium.
3. Phase transitions between condensed phases are described by the Clapeyron equation:
dp ∆H
= ,
dT T ∆V
or, in approximate form:
∆p ∆H
= .
∆T T ∆V
We calculate the right side of this equation for the ice I water transition. The volume
change is determined from the densities:
M (H 2O) M (H 2O) 18 18
∆V = V (water) − V (ice) = − = − = − 1.63 cm3/mol
ρ(water) ρ(ice) 1.000 0.917
∆p ∆H 6010 J/mol
= = −6
= − 1.35 ⋅ 107 Pa/K = − 13.5 MPa/K .
∆T T ∆V 273 K ⋅ ( −1.63 ⋅ 10 m / mol)
3
If this slope does not depend on pressure and temperature then at the pressure of 210 MPa the
temperature of liquid water in equilibrium with ice I and Ice III is approximately:
210 − 0.1
T = 273 + ∆T = 273 + = − 257.5 K = − 15.5 C .
−13.5
This is an estimate; the real value is –22 oC. The difference between the estimated and real
values is due to the fact that the enthalpy of fusion and densities vary with pressure. For
example, at 210 MPa the enthalpy of fusion of ice I is 4230 J/mol (instead of 6010 at normal
pressure), and the volume change is ∆V = –2.43 cm3/mol (instead of –1.63 cm3/mol at normal
pressure).
4. From the Clapeyron equation it follows that the slope of the p(T) dependencies for the
melting points of ice III to ice VII is determined by ∆H, T, and ∆V. The first quantity is assumed
to be the same for all transitions, the temperature is comparable in all cases, hence the main
22
contribution to the slope comes from ∆V. For ice VII, the slope is the smallest, hence, the ∆V =
V(water) – V(ice) is the largest, whereas V(ice) is the smallest. It means that ice VII is the densest
form of ice (among those forms that are shown on the phase diagram).
From the phase diagram we see that the melting point of ice VII at a pressure of 10 GPa
is about 630 K. This is, indeed, a very “hot” ice.
5. Determine the molar volume of ice VII. One mole contains NA/2 cubic unit cells:
NA 3
Vm = d = 3.01 ⋅ 1023 ⋅ (0.335 ⋅ 10−7 )3 = 11.3 cm3/mol.
2
6. Knowing the density of ice VII, we use the Clapeyron equation to estimate its enthalpy of
fusion. Comparing the triple point “water – ice VI – ice VII” and the melting point of ice VII at
pressure 10 GPa we estimate the slope: ∆p / ∆T = (104 – 2200) / (630 – 355) = 28 MPa / K. The
volume change during melting is: ∆V = (18/1.00) – 11.3 = 6.7 cm3/mol. Substituting these data
into the Clapeyron equation, we get:
∆p
∆H = T ∆V = 355 K ⋅ (6.7 ⋅ 10−6 m3 /mol) ⋅ (28 ⋅ 106 Pa/K) = 66000 J/mol.
∆T
This value is by an order of magnitude larger than the exact value 6400 J/mol. The reason is
probably due to a low resolution of the phase diagram at high pressures, which leads to a rough
estimate of the slope. This result also shows that the approximations used are not valid at high
pressures and temperatures.
lg (10−7 ) = E − 0.414
0.0591 m
E ′ = E +
m
n n
23
For this reaction, the standard Gibbs energy is 480.5 kJ/mol, and 4 electrons are transferred from
H2O to CO2. Hence, the standard emf is:
∆G 480500
E = − = − = − 1.24 V = E1 − 1.23 V
nF 4 ⋅ 96500
For CO2 reduction to carbohydrates the standard redox potential is E1 = − 0.01 V . The standard
4
biochemical potential is: E1′ = − 0.01 − 0.414 = − 0.42 V .
4
3. The overall reaction: CO2 + 2H2S → CH2O + 2S + H2O
Oxidation: H2S – 2e → S + 2H+
Reduction: CO2 + 4H+ + 4e → CH2O + H2O
Standard emf: E° = –0.01 – 0.14 = –0.15 V
Standard Gibbs energy: ∆G = − nFE = − 4 ⋅ 96500 ⋅ (−0.15) ⋅10−3 = 57.9 kJ/mol.
Energy of light with wavelength 840 nm:
hcN A 6.63 ⋅10−34 ⋅ 3.00 ⋅108 ⋅ 6.02 ⋅1023
Em = = −9
⋅10−3 = 143 kJ/mol.
λ 840 ⋅10
One quantum gives enough energy to oxidize two molecules of H2S.
24
4. Both NADP+ reduction and ATP formation require one proton, and during H2O oxidation
two protons are released. Hence, the overall reaction equation of light stages is:
NADP+ + ADP + Pi + hν → ½ O2 + NADP⋅H + ATP
(water is not present in this reaction, because the number of H2O molecules oxidized to O2 is
equal to the number of H2O molecules formed during ADP phosporylation).
6. This effect is easily understood using a simple orbital diagram (see Appendix in
“Molecular Mechanisms of Photosynthesis” by R.E.Blankenship). In the ground state, a lost
electron comes from the low-energy HOMO, while an acquired electron enters the high-energy
LUMO. As a result, the molecule is neither a strong oxidant nor a good reductant. In the excited
state, the situation is different: a lost electron leaves the high-energy LUMO, and the acquired
electron comes to low-energy HOMO: both processes are energetically favorable, and the
molecule can act both as a strong oxidant and a powerful reductant.
Ox + e → R )
(standard redox potential EOx/R
and Ox + e → R* ).
(standard redox potential EOx/R*
(
− E
F EOx/R Ox/R* = Eex = )
hcN A
λ
,
whence it follows:
= E − hcN A .
EOx/R*
λF
Ox/R
1. After the endpoint, the excessive Al3+ ions undergo hydrolysis, which makes the medium
acidic, and the indicator turns red:
Al(H2O)63+ = Al(OH)(H2O)52+ + H+
3. Cryolite Na3AlF6 being formed upon the titration is only slightly soluble in water. Hence,
NaCl was added to further decrease its solubility and shift the equilibrium of complex formation
rightwards.
6. 10.25 mL of 0.1000 M AlCl3 gives 1.025 mmol of Al3+, corresponding to 6.15 mmol of
F–. The initial amount of NaF was 0.500 g, or 11.91 mmol, i.e. 5.76 mmol of F– was spent for the
precipitation of calcium. The amount of calcium is 2.88·10–3 mol.
8. The solution of free silicic acid (a weak acid with pKa of about 10) will be slightly acidic;
hence, the indicator used in the neutralization of the sample should change its color in a weakly
acidic medium (methyl red, pKa 5). In weakly alkaline media (color change range of two other
indicators), a considerable part of the silicic acid will be present in the form of a silicate ion, the
buffer solution of which will consume a certain amount of the reacting HCl.
27
9. The amount of NaOH and the excess of HCl are the same and equal to 0.550 mmol.
Hence, the amount of HCl spent for the reaction with silicic acid is 0.994 – 0.550 = 0.444 mmol,
and the amount of silicic acid is 0.111 mmol.
3. The carbоxylic grоup cоuld either exist in the оriginal cоmpоund B (a) оr be fоrmed
during the оxidatiоn (b).
(a). Let us suppоse a minimum amount оf оxygen-cоntaining grоups in B: 0.001 mоl оf –CООH
(45 mg) and twо hydrоxyl grоups (C–ОH 29 g/mоl ∙ 0.002 mоl = 58 mg); then, 0.001 mоl оf
nitrоgen shоuld be alsо present (14 mg); this gives the total mass of 117 mg, which is even
higher than the mass оf B (105 mg). Therefоre, part оf оxygen originates frоm the оxidant оr
water as a result оf the substitutiоn оf amine nitrоgen atоm (which has transfоrmed intо the
ammоnium iоn) with оxygen (sо, aminо grоups in Malaprade reaction behave as hydrоxyl ones).
In case B cоntains one оxygen atоm less, we will get: 1 mmоl оf C–ОH grоups (29 mg) + 1
mmоl оf CHNH2 (29 mg) + 1 mmоl оf СООН (45 mg) = 103 mg. To attain the required 105 mg,
the following groups can be suggested: СНОН (30 mg), CH2NH2 (30 mg) and СООН (45 mg),
which brings us to the empirical formula of B: C3H7NО3. Remembering that nitrogen must be in
the form of an amino group, no ether oxygens are permitted and an acid is formed as the result of
28
oxidation (this can only be HCOOH, and to obtain that, a –CH(OH)– or –CH(NH2)– group must
be present), we can make up a list of possible structures of B. In case the carboxylic group was
present in the original compound (a), that will be 2-aminо-3-hydrоxyprоpiоnic acid оr 3-aminо-
2-hydrоxyprоpiоnic acid:
NH2 OH
OH OH
OH O NH2 O
(b). In case compound B is originally lacking carbоxylic grоup, the molecular weight of 105
corresponds to compounds containing 1 оxygen atоm less and 1 extra carbоn atоm (C4H11NО2),
i.e. butane derivatives containing 1 amino and 2 hydroxyl groups. (Propanes with 3 hydroxyl
groups and 1 amino group will give molecular weights of 107.) If the butane moiety is
unbranched and all three grоups (twо ОН and the NH2) are vicinal:
OH OH NH2
NH2 OH OH
then periоdate оxidatiоn yields HCHО, HCООH and CH3CHО; the two isоbutane derivatives
HO H2N
OH OH
NH2 OH
H3C H3C
29
both result in 1 mol of CH3COOH and 2 mоles оf HCHО, while all these butanes meet the task
requirements, as they require 8 mmol equivalents оf KMnО4 for their oxidation. Compound
HC(CH2ОH)2CH2NH2 is nоt оxidized with periоdate. If ОН and NH2 groups are nоt vicinal,
fоrmaldehyde and an aldehyde are fоrmed, but no necessary carbоxylic acid is produced. If B
cоntains a C=О group (for instance, O=CH–CH(OH)–CH(NH2)–CH3), its fоrmula is C4H9NО2
(mоlecular weight 103), which is not consistent with the problem conditions.
Scheme of periodate oxidation for the linear butane derivatives:
CH2NH2–CH(OH)–CH(OH)–CH3 CH2O + HCOOH + CH3CHO
For the branched butanes:
HОCH2–C(CH3)NH2–CH2ОH CH2O + CH3COOH + CH2O
4. The amount of chromium is found as follows: 3nCr = nthios = 0.0485 M·5.01 mL = 0.2430
mmol (nCr = 0.081 mmol). This corresponds to 26.2 mg of PbCrO4 (M = 323.2 g/mol) in the
aliquot, or 131 mg totally.
7. On acidification of the 2nd aliquot, chromium is reduced by [Fe(CN)6]4– (see i. 5). Then
permanganate is spent for the oxidation of [Fe(CN)6]4–, namely, the amount of [Fe(CN)6]4– added
plus the amount contained initially in the sample less the amount spent for the reduction of
Cr(VI): 5nMnO4 = nFe added + nFe from sample – 3nCr. From this equation we can find nFe from sample =
5nMnO4 – nFe added + 3nCr = 5·0.00500·2.85 – 10·0.0300 + 0.2430 = 0.07125 – 0.3000 + 0.2430 =
0.0142 mmol [Fe(CN)6]4–. One mol of [Fe(CN)6]4– results from 0.5 mol of Fe3[Fe(CN)6]2 (see
solution to question 1), therefore, the answer is 4.21 mg of Fe3/2[Fe(CN)6] (M = 295.6 g/mol) in
the aliquot, or the total amount of 21.1 mg.
The interaction between A and B under acidic condition proceeds as Friedel-Crafts alkylation of
the aromatic ring with the thermodynamically more stable secondary propyl carbocation as an
electrophile. Being a product of the interaction of equal amounts of A and B, C turns out to be
31
isopropylbenzene, i.e. cumene. Oxidation of C with subsequent acidification leads to phenol and
acetone D. This classical industrial procedure is known as cumene process.
The structure of D can also be easily determined from that of bisphenol A, which is formed as a
result of two consecutive Friedel-Crafts alkylations of phenol. Treatment of bisphenol A with
NaOH leads to disodium bis-phenolate E, which gives polycarbonate with the monomeric unit F
as a result of the reaction with phosgene.
The reaction of phenol with diluted nitric acid proceeds as a mononitration resulting in isomeric
nitrophenols G and H. Due to the activation effect of OH-group in phenol, electrophilic
substitution can occur in ortho- and para-positions of phenol. G is para-nitrophenol (two planes
of symmetry), whereas H is ortho-nitrophenol (only one plane of symmetry). Further reduction
of NO2-group in para-nitrophenol G results in para-aminophenol I. Due to its higher
nucleophilicity, NH2-group (rather than OH-group) in I is acetylated with acetic anhydride
giving paracetamol J.
32
The reaction of phenol with CO2 in the presence of NaOH proceeds through intermediate
formation of sodium phenolate, which interacts with CO2 under heating and high pressure
(Kolbe-Schmitt reaction) to give disodium salicylate K. Acidification of K with two equivalents
of an acid results in salicylic acid L, which provides aspirin M when acetylated with acetic
anhydride.
Aluminon synthesis is based on the same approach as previously considered for bisphenol A. The
reaction of salicylic acid L with formaldehyde under acidic conditions affords N, which is an
analogue of bisphenol A. Addition of another equivalent of salicylic acid L under oxidative
conditions (NaNO2/H2SO4) gives the tri-acid O, which is a direct precursor of Aluminon. Thus,
the structure of O can be derived from that of Aluminon.
33
Problem 19. Chrysanthemic acid
1. Chrysanthemic acid is formed as a result of hydrolysis of its ethyl ester, F, which, in turn,
is obtained by cyclopropanation of E with diazoacetic ester. Therefore, E is 2,5-dimethylhex-
2,4-diene with molecular formula C8H14. This conclusion is supported by the molecular formula
of D. Evidently, transformation of D to E is elimination of two water molecules.
Eight carbon atoms of D originate from A and B. The other reaction between these compounds
affords L containing 5 carbon atoms (N is formed from H and C7H7O2SNa; the number of
carbon atoms in H and L is the same). The provided information strongly suggests that A is
acetylene (C2H2). Hence, A is composed of 2, and B should be composed of 3 carbon atoms.
Reaction between A and B was disclosed by Favorskii in 1905 as that between acetylenes and
carbonyl compounds. It means that B is either propionic aldehyde (C2H5CHO) or acetone
(CH3COCH3). Accounting for the structure of E, B is acetone. E is also formed through the
Grignard reaction of acetone with the corresponding RMgBr followed by elimination of water.
The structure of H can be unambiguously deduced from that of E – it is prenyl bromide. So, the
natural alcohol is prenol (3-methylbut-2-en-1-ol). L is formed from A and B under the same
reaction conditions but when A to B ratio is of 1:1. Therefore, L is 2-methylbut-3-yn-2-ol. Its
hydrogenation in the presence of Lindlar catalyst leads to the corresponding alkene M.
Subsequent reaction with HBr affords prenyl bromide H via nucleophilic substitution with
double bond migration. The reaction of H with sodium 4-toluenesulfinate results in the
corresponding sulfone N.
O
B H2 HO
C2H2 HO OH
KOH, C6H6 Pd/C OH
A 0,4-0,9 MPa C D
H+
t
1) Mg
PBr3 2) (CH3)2CO N2CHCO2C2H5 OC2H5
OH Br 3) H+, t
O
G H E F'
O
O
SO2Na
B H2 HBr S
C2H2 HO HO O
Br
KOH, C6H6 Pd/BaSO4
0,4-0,9 MPa quinoline H
A L M
N
34
Finally, acid-catalyzed self-condensation of acetone yields 4-methylpent-3-en-2-one (mesityl
oxide, I). Iodoformic reaction of I produces the salt of the corresponding acid J which is further
transformed into ethyl ester K. The reaction of K with deprotonated sulfone N results in
chrysanthemic acid ester F”.
O O O O
H3O+ I2 EtOH N
OC2H5
NaOH ONa H+ OC2H5 NaOCH3
O
B I J K F''
2.
3. The first step is the Diels-Alder reaction. Compound P with tetrasubstituted double bond
is the most stable isomer of O with the same carbocyclic skeleton. Heating of P with ammonia
leads to imide R, which further reacts with CH2O giving the target alcohol X.
O O O O O
t Pd/C NH3 CH2O
+ O O O NH N
t OH
O O O O O
O P R X
4. Amides and hydrazides do not easily form esters in reaction with alcohols. Oppositely,
anhydrides are appropriate reagents for the ester synthesis. Moreover, re-esterification of methyl
or ethyl esters with high-boiling alcohols is well-known. These reactions are efficient due to
methanol (ethanol) removal from the reaction mixture via distillation (Le Chatelier’s principle).
Molecular formulae of the esters formed from alcohol S and T are C21H20Hal2O3 and
C22H19Hal2NO3, respectively. Halide content in the esters is 2MHal/(2MHal + 320) and
2MHal/(2MHal + 345), respectively. Calculation of halide content in these compounds allows
unambiguously deciding on the structures of the pyrethroids.
Content of Hal, % Using exact atomic mass Using approximate atomic mass
F Cl Br I F Cl Br I
X R2 R2
R2 X
X R1 R1
R1 H+ H+
NH2 +
N N NH
H O N N
H H
A
3. Reactions are started by interaction of amine with the carbonyl group furnishing imine.
To complete pyrrole moiety formation, monoimine of hexane-2,5-dione should isomerize into
the enamine followed by an attack of the amine group on the second C=O group. Formally,
imine of pentane-2,4-dione can form the pyrrole ring in two ways. First, it is the interaction of
the nitrogen atom with the methyl group. However, the methyl group itself is unreactive towards
nucleophiles. Keto-enol equilibrium with involvement of the methyl group in this compound is
less probable than that with CH2-fragment. Even if the equilibrium was true, enol is a
nucleophile and cannot react with nucleophilic nitrogen atom. Therefore, the second possibility
should be considered, namely, the reaction of the second carbonyl with CH2 bound to N atom.
This reaction is quite probable as CH2-group is also connected with the electron-withdrawing
ester group and can be deprotonated by a base as shown below.
O
H3 C OC2H5 H+
+ H N H3C CH3
CH3 2 N
O O
CO2C2H5
B
CH3
H 3C CH3 OC2H5 H3C CH3 base
+ H N
2 H3C N CO2C2H5
O O O N O H
CO2C2H5 C
37
4-5. Two products are formed in the reaction of propyne, and only one product in the case of
the alkyne bearing an electron-withdrawing ester group. This allows supposing a nucleophilic
attack of a certain intermediate on the alkyne moiety. A base generates a nucleophilic agent from
acetone oxime. Again, two ways of deprotonation are possible: O-deprotonation and C-
deprotonation. However, oxime enolate, if formed, should add to alkyne with the formation of
hex-4-en-2-one oxime. There is no possibility for the transformation of this oxime into pyrrole
ring. The alternative possibility is O-deprotonation and nucleophilic addition of the oximate ion
to alkyne furnishing O-alkenyl acetone oxime. Formation of the C-C bond between the methyl
group of acetone and the -carbon atom of the alkenyl group is needed to complete the pyrrole
ring synthesis. At the first glance, such transformation is impossible. However, this system is
very similar to the N-aryl-N’-alkenyl moiety which undergoes the 3,3-sigmatropic rearrangement
in the Fischer indole synthesis. Indeed, isomerization of O-alkenyl acetone oxime into O-
alkenyl-N-alkenyl derivative creates the fragment required for the 3,3-sigmatropic shift. So,
formation of the pyrrole ring giving 2,4-dimethylpyrrole and 2,5-dimethylpyrrole is analogous to
that of indole in the Fischer synthesis. The former compound is transformed into C via N-
deprotonation followed by the Kolbe-Schmitt carboxylation and ester formation. To provide B,
E should be N-alkylated with ethyl haloacetate. Halogen can be determined from the carbon
content in the alkylation reagent.
38
CO2C2H5
CO2C2H5
DMAP
+
PhMe N
N microwaves H
OH
G
6. Methyl group in the starting compound is very acidic due to activation by both ortho-
nitro group and para-nitrogen atom of pyridine. So, it can be easily deprotonated to further react
with diethyl oxalate providing the corresponding ketoester H. Reduction of the nitro group gives
aniline. Condensation of the amino group with the appropriately located ketone moiety affords
the 6-azaindole derivative I (C11H12N2O3). Aminomethylation of this indole furnishes the
gramine derivative J which undergoes nucleophilic substitution with sodium dimethylmalonate
producing K. Its hydrolysis results in a compound with the molecular formula of C 11H10N2O5. It
means that: a) hydrolysis of the malonate fragment is accompanied by decarboxylation; b) the
ester moiety at the C2 position of the indole is hydrolyzed too. However, even if so, the
molecular formula should be C12H14N2O5. The difference equals to CH2. Hydrolysis of OCH3-
group in ortho-position to pyridine nitrogen is the only possibility. Indeed, hydrogenation of this
pyrrolopyridone yields M. Its decarboxylation and hydrolysis of the amide function finally leads
to porphobilinogen.
39
Problem 21. Cyclobutanes
1. Hydrocarbon K consists of 90% C and 10% H. Its simplest formula is (C3H4)n, and it has
a single type of H atoms. So, it is allene, H2C=C=CH2. A has 5 carbon atoms. Therefore, allene
reacted with CH2=CHCN in a ratio of 1:1 and lost one carbon atom during the following steps.
Various products can be supposed for this reaction, however, it is known that allene is prone to
undergo cycloaddition as 2-component. Acrylonitrile undergoes cycloaddition as 2-
component too. So, the product should be a cyclobutane derivative, which is consistent with the
next scheme. The C and H content in N and O provides for their molecular formula
(C5H10Br2O2). In this respect, two sub-processes should proceed: a) acetone is doubly
brominated; b) ketone is transformed into ketal in the reaction with methanol catalyzed by HBr
evolved in the bromination step. Two dibromoacetones (1,1- and 1,3-) can be formed. Reaction
of the latter with dimethyl malonate affords the corresponding cyclobutane derivative. Its
treatment with hydrochloric acid leads to the hydrolysis of ketal into ketone and esters into an
acid. So, the product should be 3-oxocyclobutane-1,1-dicarboxylic acid. However, its formula is
C6H6O5. Therefore, hydrolysis is also accompanied by decarboxylation of the malonic acid
moiety. So, A is 3-oxocyclobutanecarboxylic acid. Accounting for it, L is the product of [2+2]
cycloaddition, i.e., 1-cyano-3-methylenecyclobutane. Its hydrolysis followed by oxidation of
C=C double bond produces A. Finally, the schemes for preparation of A are as follows:
MeO OMe O
O Br2 MeO OMe
MeO OMe CH2(CO2Me)2 20% HCl
Br + Br Br
CH3OH t
Br
rt MeO2C CO2Me CO2H
N O P A
Reaction of A with SOCl2 furnishes acyl chloride, which reacts with NaN3 affording acyl azide.
Heating of RCON3 produces isocyanate R-N=C=O, which immediately reacts with t-BuOH
giving rise to N-Boc-protected 3-aminocyclobutanone B. Reduction of keto group with NaBH4
leads to cis- and trans-isomers of the corresponding aminocyclobutanol. Further reaction with
CH3SO2Cl produces mesylates, which undergo SN2 displacement with NaN3 affording
aminoazides. Reduction of azido group and deprotection of amine furnishes cis-and trans-
isomers of 1,3-diaminocyclobutane. Therefore, J is cis-1,3-diaminocyclobutane (two planes of
symmetry), and I is trans-isomer (one plane of symmetry). Similarly, G is trans-, and H is cis-
40
isomer. As the SN2 reaction proceeds with inversion of the configuration, compound E (leading
to G) is cis-, and F is trans-isomer.
O OH OH
O O
O
1) SOCl2; t-BuOH NaBH4 +
OH OSO2CH3 N3 NH2
NaN3 1) H2, Pd/C
CH3SO2Cl
2) CF3CO2H
Et3N 3) NaHCO3
HN OCMe3 HN OCMe3 HN OCMe3 NH2
O O O
C E G I
OH OSO2CH3 N3 NH2
NaN3 1) H2, Pd/C
CH3SO2Cl
2) CF3CO2H
Et3N 3) NaHCO3
HN OCMe3 HN OCMe3 HN OCMe3 NH2
O O O
D F H J
2-3. Reduction of P with LiAlH4 gives the corresponding diol Q, which is transformed into
ditosylate R. Reaction of R with dimethyl malonate leads to formation of the second cyclobutane
ring (S). Hydrolysis of S proceeds similarly to that of P, i.e., it produces ketoacid T. Further
transformations are also similar to those in the first scheme and produce spiro[3.3]heptane-2,6-
diamine W. This compound has no plane or center of symmetry. It is chiral due to axial chirality
(similarly to 1,3-disubstituted allenes), thus, it can be resolved into two enantiomers.
MeO OMe
MeO OMe MeO OMe MeO OMe
1) LiAlH4 TsCl CH2(CO2Me)2 20% HCl
2) H2O Py t
HO OH TsO OTs
MeO2C CO2Me MeO2C CO2Me
P Q R S
O NOH
O H NH2
1) H2, Pd/C
1) SOCl2; 2) NaN3 NH2OH 2) CF3CO2H
3) NaHCO3
3) , t-BuOH
HN OCMe3 HN OCMe3
CO2H NH2
T U O V O W
41
Problem 22. Introduction to translation
2. No, because of redundancy of the genetic code: most amino acids are encoded by several
codons.
O O O
O A
O P O P O P O
H3N CH C O
+ O CH2
O O O
R H H (1)
H H
OH OH
O O
A
H3N CH C O P O
O O
OH CH2
R O
H H + O P O P O
aminoacyl adenylate H H O O
OH OH
42
C
O C
H H
H H A O
O OH
H H
O P O H H
O O OH A
- H H O
O
O P O + -
A
H H O O P O (2)
OH OH -
O H H CH2
OH O
O O H H H H
H2N CH C O P O
+ A O OH
H H
CH2 O C OH OH
R OH O
H H CH R
H H NH2
OH OH
aminoacyl-tRNA
Thus, the carboxylic group of the amino acid reacts with 3’-OH group of its tRNA.
5. а) Met-Asp-His-Ala-Ile-Asn-Val-Val-Gly-Trp-Ser-Val-Asp-Thr-Leu-Asp-Asp-Gly-Thr-
Glu-Ala or fMet-Asp-His-Ala-Ile-Asn-Val-Val-Gly-Trp-Ser-Val-Asp-Thr-Leu-Asp-Asp-Gly-
Thr-Glu-Ala, depending on the biosynthesizing species (Eukaryotes, Prokaryotes, or Archaea).
b) The third amino acid is tyrosine, and the last one is valine. All the rest positions are the same.
c) The N-terminal amino acid is leucine. All the rest positions are the same. It should be noted
that the translation in bacteria would not start without the START codon.
d) The last but one codon is changed into STOP codon, which will result in the oligopeptide
shorter by 2 amino acid residues than that in i. 5a.
6. AUG-GAU/C-GUN-AAU/C-CAU/C-CCN-GAA/G-UAU/C-GGN-AAA/G
8. Taking into account that the A:C ratio is 1:5, the probability of finding A and C at any
position is 1/6 and 5/6, respectively. Thus, the probability of finding certain codons is:
AAA =(1/6)3=1/216 CCC=(5/6)3=125/216
AAC=(1/6)2*5/6=5/216 CCA=(5/6)2*1/6=25/216
ACA=1/6*5/6*1/6=5/216 CAC=5/6*1/6*5/6=25/216
ACC=1/6*(5/6)2=25/216 CAA=5/6*(1/6)2=5/216
Using the table of genetic code one gets: Lys:Asn:Thr:Pro:His:Gln=1:5:30:150:25:5
43
9. Anticodon has no influence on the CCA3’ terminus. Thus, the mutant tRNA will add
tyrosine to the positions where serine was initially expected with respect to mRNA sequence.
This may lead to improper folding of the protein and total or partial loss of its functional activity.
10. Glu is encoded by GAA and GAG, and His by CAU and CAC. Two substitutions (of the
1st and 3rd residues) are needed to make this mutation true, which is quite improbable. Single
residue mutations occur much more frequently, and Glu to Gln mutation can serve as an example
(together with many other mutations of this type).
2. Both α-carboxylic and α-amino groups exist mostly in the ionic forms at pH 4.7.
Ionization state of the side groups at the given pH value should be determined individually based
on their pKa values as reported in Wikipedia. One should leave into consideration only amino
acids with the number of atoms less than 19 (28-10=18; this is maximal possible value in case
one of two amino acids is glycine). Surprisingly, the data found on different Wikipedia pages
lead to contradictory results. According to the former weblink
(http://en.wikipedia.org/wiki/Proteinogenic_amino_acid), only ten amino acids can be further
considered. These are:
44
H COO- - COO-
Gly 10 Asp OOC 15
NH3+ NH3+
COO- COO-
Ala 13 Pro +
N H 17
NH3+
H
COO- COO-
Cys HS 14 Thr HO 17
NH3+ NH3+
+
COO- H3N COO-
Sec HSe 14 Asn 18
NH3+ O NH3+
COO- -
OOC COO-
Ser HO 14 Glu 18
NH3+ NH3+
The listed amino acids provide for the following dipeptides (without regard to N- and C-termini):
Ser-Cys, Ser-Sec, Cys-Sec, Gly-Asn, Gly-Glu и Asp-Ala. Taking into account the residue
positioning (N- or C-terminal), one gets two different dipeptides for each of 4 former pairs, and 3
dipeptides for each 2 latter pairs (note that β-carboxyl group of Asp and γ-carboxyl group of Glu
can be also involved in peptide bond formation; see an example below).
O COO-
O NH3+ O
-
OOC N COO- N COO- -
OOC N COO-
+H +H H
NH3 NH3
(1) (2) (3)
Thus, the total number of dipeptides equals to 14. However, serious caution is needed when
using Wikipedia, since it is a collection of the user-generated content. Note that pKa values of
some groups are absolutely incorrect (section “Side Chain Properties”). In particular, the side
group of Asn is absolutely non-protonated at pH 4.7. Finally, the correct number of individual
peptides is 12 (excluding Gly-Asn and Asn-Gly).
45
Screenshot of the webpage http://en.wikipedia.org/wiki/Proteinogenic_amino_acid dated
20.10.2012 is given below. Being irresponsible of these mistakes, authors of the problem
promise to correct the data after publishing the Solutions to Preparatory problems.
At the same time, the pKa values found at the other webpage
(http://en.wikipedia.org/wiki/Amino_acid) are correct.
3. One should analyze all five variants of dipeptides (with no regard to N- and C-termini)
from i. 2 by calculating masses of corresponding precipitates. Typical procedure is given below
for the correct answer (Cys-Sec):
C6H12N2O3SSe + 9.5O2 → 6CO2 + SO2 + SeO2 + N2 + 6H2O (1);
Ca(OH)2 + CO2 → CaCO3↓ + H2O (2);
Ca(OH)2 + SO2 → CaSO3↓ + H2O (3);
Ca(OH)2 + SeO2 → CaSeO3↓ + H2O (4).
Number of moles of dipeptide: 1.000 g 271.19 g/mol = 3.687 ∙ 10−3 mol. Thus, the mass of
precipitate is:
m precipitate = 3.687 ∙ 10−3 mol ∗ 6 ∗ 100.09 + 120.14 + 167.04 g/mol = 3.273 g
However, further calculations according to the equations of chemical reactions of precipitate
dissolution in hydrochloric acid
CaCO3 + 2HCl → CaCl2 + CO2↑+ H2O (5);
CaSO3 + 2HCl → CaCl2 + SO2↑+ H2O (6),
provide for contradictory results. Gas volume given in the task is by approx. 15% less than that
obtained from the calculations. The only reason behind the difference is the deficiency of
46
hydrochloric acid with respect to the precipitate amount (Note that by contrast to the rest of the
task, there is no indication of an excess or deficiency in the case of hydrochloric acid!).
Since the available data is insufficient to decide on the sequence of amino acid residues, both
Cys-Sec and Sec-Cys are accepted as correct answers for X.
SeH SH
O O
(R) (R) (R) (R)
HS N COOH HSe N COOH
H H
NH2 NH2
(1) (2)
4. The –SeH group is a much stronger reducing agent than the –SH group. Thus, Sec is very
readily oxidized, which makes its presence as free selenocysteine inside a cell impossible.
5. Searching for a correlation between the given images and sequences is much easier than
can be expected. There could be many ways to reach the correct answer. A sample strategy is
given below. First, one should decide which of the fragments refers to human RNA. Genomes of
the viruses belonging to the same family should be phylogenetically close, with a slight
divergence form the common ancestor. Indeed, sequences 1 and 3 reveal high similarity, both
dramatically differing from sequence 2, the latter thus being attributed to human cell. Next step
is the search for nucleotides corresponding to the black boxes in the image of human RNA. Note
that there are colorless and grey boxes to the ends from black ones. These include 9 nucleotides
at the 5’- and 11 nucleotides at the 3’-end. These forbidden areas are highlighted red in the
hereunder sequence. Nucleotides corresponding to the black boxes are located between the red
fragments, and should be twice two consecutive purine nucleotides AG, AA or GG (all options
highlighted yellow). Furthermore, there should be exactly 30 nucleotides between the yellow
fragments, which allows the final assignment (highlighted yellow and underlined).
AGGCACUCAUGACGGCCUGCCUGCAAACCUGCUGGUGGGGCAGACCCGAAAAUCCCAC
Thus, the encircled codon is UGA, which can be also found in fragments 1 and 3. Using the
above strategy, one can fill in the rest two images and find the correlation between the images
and fragments (fragments 1, 2, and 3 refer to the images of the fowlpox virus, homo sapiens, and
canarypox virus, respectively).
47
N H O N
N
R O H N N
R
N
H2N
7. UGA, according to the table of genetic code, is known as the STOP codon terminating
translation. However, it is stated in the problem that the chain elongation proceeds after UGA
(variants 2 and 3 invalid). UGA is similarly located in sequences of very dissimilar organisms (a
mammal and viruses), which underlines its importance for translation and makes variant 5 hardly
possible. Variant 4 can be also discriminated, since translation is an uninterruptible process.
Thus, variant 1 is the correct answer. Indeed, UGA in a certain motive (referred to as SECIS
element, Selenocysteine Insertion Sequence) is read as the codon determining selenocysteine
inclusion into polypeptides. In viruses, SECIS element is located in the translated region of
RNA. In eukaryotes, this hairpin-like structure is found in the unreadable part of mRNA (in 3’-
untranslated region, 3’-UTR), and Sec is not found in human proteins.
8. Knowledge of the UGA position allows setting the reading frame. In principle, there
could be various mutations meeting the requirements. Examples are given below.
48
Choosing a mutation, one should keep in mind that the wild type and mutant codons must encode
the same amino acid. Also, nucleotides of this codon should not be involved in maintaining the
secondary structure of SECIS element (no hydrogen bonding to opposite nucleotides). Thus, one
can suggest U-23→С-23 mutation for the fowlpox virus (both are tyrosine codons), and A-
28→G-28 mutation for the canarypox virus (both are lysine codons).
1. Calculation of molar ratios of carbon, hydrogen and oxygen in A-C allows determining their
minimal molecular weights corresponding to the net formulae (note that isotopic ratios of C, H,
N and O are native):
Calculation of minimal
Compound Calculation of ratios
molecular weights, g/mole
31.09 5.74 16.57 60.05 ∙ 100
A 𝑛 C ∶ 𝑛 H ∶ 𝑛(O) = : : = 5 ∶ 11 ∶ 2 𝑀= = 193.1
12.01 1.008 16.00 31.09
With provision of the upper bound (M<250 g/mole), true and minimal molecular weights
coincide. The residual molecular weights available for the other two elements (besides C, H, and
O) in A and B are of 90.0 and 91.0 g/mole, respectively. There are two possible reasons behind
the difference in the residual molecular weights for A and B (91.0-90.0=1 g/mole). These are
dissimilarity of atomic weights of the fifth elements in A and B and/or different number of
nitrogen atoms in these compounds. All possible variants of the number of nitrogen atoms
(cannot exceed 6) in A are considered in the hereunder table:
Number of N Residual molecular weight left Variants of the 5th
Biochemical sense
atoms in A for the 5th element in A element
1 76 - -
2 62 2P To be considered
1 Ti? Impossible
2 Mg? Impossible
3 48
3 O? Impossible
4 C? Impossible
4 34 - -
5 20 1 Ne? Impossible
6 6 6 H? Impossible
49
With provision of the inequality given in the problem text, the variant of 2 nitrogen atoms in A
corresponds to 1 or 2 nitrogen atoms in B, and 75 or 63 g/mole left for the 5th element in the
latter compound, respectively. No reasonable variants are in agreement with the above values.
Therefore, we seem to have come up against a brick wall.
2. Difference by 1 g/mole in the molecular weights of the 5th element in A and B is left as the
only reason. This can be true in case of isotopes (note that native isotope ratios are mentioned
only for four elements!). If so, isotopes should be stable (stability of all initial compounds) and
most likely of one and the same element (A, B, and C are precursors of the same compound X).
With account of the equal number of nitrogen atoms in A and B, the following set of isotope
combinations is available: 20-21, 34-35, 48-49, 62-63, 76-77. Furthermore, the difference of 1
g/mole unambiguously suggests only one atom of the 5th element in each of A and B.
Two sets of stable isotopes (48Ti-49Ti and 76Se-77Se) formally fit well. Since there are no native
titanium-containing amino acids, the elemental composition of A and B is finally found as: C, H,
N, O, and Se.
4. Both R- and S-amino acids are found in nature. Since it is not mentioned in the problem
text which exactly of A and B is found in proteins, it is impossible to unambiguously assign
configurations of α-carbon atoms without additional information.
5. Gases A1, B1, and С1 have molecular weights of 106, 107 and 112 g/mole, respectively. It
is seen that the difference in the molecular weights of A and B (1 g/mole) is retained for their
metabolites. Thus, A1 and B1 are likely to be isotopologues. Besides selenium, A1 and B1
contain elements with a total residual molecular weight of 106 – 76 = 30 g/mole. Since gaseous
metabolites contain hydrogen, there are two possible variants of their molecular formula: C2H6Se
or CH2SeO. With provision of identity of hydrogen atoms in A1, the following structures are
possible:
H3C CH или
or H2C Se
Se 3
O
50
Of these two, only dimethylselenide does not contain π-bonds. Finally, A1 is (CH3)276Se, and B1
is (CH3)277Se.
6. The atomic weight of selenium isotope in С1 is 76 + (112-106) = 82 a.u. (Note that the
final metabolite is the same for all three initial original compounds!). Residual molecular weight
left for the 4th element in C (it consists of only four elements) is 130 – 16 – 82 = 32 g/mole,
which corresponds to two atoms of oxygen. Thus, the molecular formula of C is CH4O282Se.
Presence of methyl groups in C1 as well as lack of C–O bonds in the structure allow the final
ascertainment of the structural formula of C (the leftmost of the hereunder ones with 82Se):
H3C OH O H
Se или
or H3C Se
O O
NH2 NH2
селенометионин
Selenomethionine метионин
Methionine
76
Isotope Se is found in nature (~1% of the total selenium pool), so the residue of
selenomethionine with 76Se can be found (though rarely) in proteins.
NH2 NH2
Variants 3 and 5 are impossible, since protein biosynthesis admits the only way of polypeptide
chain elongation, which involves an amino acid residue transfer from aminoacyl-tRNA.
COOH COOH
Se S
NH2 NH2
метилселеноцистеин метилцистеин
Methylselenocystein Methylcysteine
e
1. Glucose consists of carbon, oxygen and hydrogen. As a result of its fermentation in H2O
the following gaseous (at STP) products could be theoretically formed:
1) Molecular hydrogen,
2) Various hydrocarbons,
3) Formaldehyde,
4) CO and CO2.
Absence of C-H bonds in C and D allows excluding variants 2 and 3 from further consideration.
Molecular mass of the gas mixture is 10.55*2 g/mol = 21.1 g/mol. It is obvious that hydrogen is
one of the two gases, whereas either CO or CO2 is the other one. CO seems to be an improbable
variant; still all the options should be checked by applying the hereunder formula for n:
52
𝑛 10
𝑀 𝑪 ∙ +𝑀 𝑫 ∙ = 21.1
𝑛 + 10 𝑛 + 10
211 − 10 ∙ 𝑚(𝑫)
𝑛=
𝑀 𝑪 − 21.1
C D Coefficient n
H2 CO2 12.0
CO2 H2 8.3
H2 CO 9.6
CO H2 27.7
Since n is integer in only one case, C and D are attributed to H2 and CO2, respectively.
Note that bacterial cultures exist in specific, sometimes solid, nutritious media. Thus,
conventional data of gases (in particular, of CO2) solubility in water may be inapplicable.
2. With respect to the results in i. 1, the updated reaction (1) is rewritten as:
5С6H12O6 + kH2O → lA + mB + 12H2 + 10CO2
a) In the case when each of A and B is a saturated monocarboxylic acids, the equation
transforms into:
5С6H12O6 + kH2O → lСxH2xO2 + mСyH2yO2 + 12H2 + 10CO2,
where x and y are the numbers of carbon and hydrogen atoms in C and D, respectively.
With account of the balance of the elements numbers, one gets the hereunder system of
equations:
It is seen from the first two equations that k = 2. Thus, the equation for oxygen can be rewritten
as l + m = 6
A and B with higher number of carboxylic groups (for example, two dicarboxylic acids) are
impossible, as this results in negative k, l, or m.
3. l and m are integers, and l + m = 6. This suggests the following possible ratios: 1:1 (3:3),
1:2 (2:4) and 1:5. Still, l·x +m·y = 20, which makes the ratio of 1:1 impossible (both x and y non-
integer, 20/3 = 6.67). Ratios of 2:1 and 5:1 are theoretically possible. Thus, the correct variants
are b, e and f.
4. The next step is a search for integer solutions of the equation l·x + m·y = 20 for the ratios
established in i. 3.
l = 2; m = 4 l = 1; m = 5
х y х y
8 1 15 1
6 2 10 2
4 3 5 3
2 4
Since the number of carbon atoms decreases as a result of fermentation (x<6 and y<6), only the
variants underlined in the above table are left for consideration. These correspond to four
unbranched monocarboxylic acids:
H3C COOH COOH COOH COOH
acetic acid propanoic acid butyric acid valeric acid
The given limitation of less than 100 atoms in each of Zstart and Zfinish can be written as n<6.
Variable b is necessarily integer, thus leading to the solely possible combination of b=55 and
n=3. So, Zstart and Zfinish are composed of 58 and 55 atoms, respectively. This means that Zstart
loses 3 atoms in acetyl-CoA formation.
54
6. The difference in the number of hydrogen atoms in Zstart and Zfinish is:
Δ𝑁𝐻 = 𝑁𝐻 𝒁𝐬𝐭𝐚𝐫𝐭 − 𝑁𝐻 𝒁𝐟𝐢𝐧𝐢𝐬𝐡 = 58 ∙ 0.43103 − 55 ∙ 0.41818 = 2
Thereby, two of three atoms appearing in acetyl-CoA from Zstart are hydrogen atoms. Oxygen or
carbon can be the third atom lost by Zstart. In the former case, Zstart loses H2O, and in the latter
case CH2-group, which is formally equivalent to substituting a CH3-group with 1 hydrogen atom.
Equation (5) is invalid with any E, whereas equation (4) is correct, if E is carbon monoxide CO
formed via enzymatic reduction of CO2.
Since bacteria cultivation proceeds in the presence of isotope-labeled CO2, the number and
isotope distribution of nitrogen atoms in acetyl CoA are not influenced. Thus, the molecular
mass of acetyl-CoA isotopologues is:
100 ∙ 14.01 ∙ 7
𝑀 𝑙𝑎𝑏𝑒𝑙𝑒𝑑 𝐴𝑐𝐶𝑜𝐴 = = 811.8 𝑔/𝑚𝑜𝑙
12.08
Molecular mass of unlabeled acetyl-CoA is 809.5. With account of rounding of nitrogen mass
fractions, the difference is of 2 g/mol. Two hereunder variants are possible:
1) CO2 labeled with 13C enters the reaction, thus giving acetyl residue with two 13C atoms;
2) CO2 labeled with two 18O enters the reaction, thus giving acetyl residue with 18O atom.
7. The initial nucleotide ratio is 1:1, thus the probability of finding G or C at any position
equals ½. Hence, the probability of any of eight possible codons is of 1/2*1/2*1/2=1/8. Four
amino acids are each encoded by two codons composed of only G and/or C. Thus, the ratio
55
between Pro, Arg, Gly, and Ala is 1:1:1:1. However, with account of the limited length of the
oligopeptide (about 33 amino acid residues), there could be significant deviations from the above
ratio. So, variant 6 is the most correct choice.
8. Using the table of genetic code, one can write down the nucleotide sequences of the
initial and mutant mRNA fragments (see designations of N and N1/N2 in Problem 22, i. 6):
UGG-CAU/C-AUG-GAA/G-UAU/C (initial);
UGG-ACN-UAU/C-GGN-GUN (mutant).
Comparison of two sequences suggests that the mutation (insertion of A) occurred right after the
first codon. Mutations influencing polypeptide biosynthesis are classified into two groups: the
substitution of base pairs and the frameshift. The latter happens upon deletion or insertion of
nucleotides in a number not multiple of three. Then, the initial sequence can be rewritten as:
UGGCAUAUGGAGUAU/С
If the mutant protein ends up with the 234rd amino acid residue, the biosynthesis is terminated by
a STOP codon present next. Since STOP codons always start with U, the completely deciphered
sequence of nucleotides is:
UGGCAUAUGGAGUAU
[𝐀𝐛∗𝐀𝐠]
1. K𝑏 = 𝐀𝐛 ∙[𝐀𝐠]
Kb = 1010
1.0
Kb = 109
0.8
Kb = 108
0.6
n
0.4
0.2
0.0
0.0 2.0x10-8 4.0x10-8 6.0x10-8
[Ag]
56
As seen, the titration curves are strongly non-linear, which makes their analysis complicated.
[𝐀𝐛∗𝐀𝐠] [𝐀𝐛∗𝐀𝐠]
3. K𝑏 = 𝐀𝐛 ∙[𝐀𝐠]
⇒ [𝐀𝐠]
= K 𝑏 (C𝐀𝐛 − [𝐀𝐛 ∗ 𝐀𝐠])
Thus, a plot in such coordinates (referred to as the Scatchard ones) should be a straight line with
the slope of –Kb and the intercept of CAbKb (CAb is the total Ab concentration).
1.5
[Ab*Ag]/[Ag]
1.0
0.5
0.0
0 20 40 60 80
[Ab*Ag], mol/L
20
15
[Ab*Ag]/[Ag]
10
0
0 20 40 60 80
[Ab*Ag], mol/L
4. If all the binding sites are independent, and Kb does not depend on the fraction of
occupied binding sites, mathematically there is no difference between x molecules of antibody
with valence N and N·x molecules of antibody with valence 1. Thus, the above mentioned
Scatchard equation is only slightly modified to account for several binding sites per antibody:
[𝐀𝐛∗𝐀𝐠]
[𝐀𝐠]
= K 𝑏 (N · C𝐀𝐛 − [𝐀𝐛 ∗ 𝐀𝐠]).
57
1.4
1.2
1.0
[Ab*Ag]/[Ag]
0.8
0.6
0.4
0.2
2 4 6 8 10 12 14 16 18
[Ab*Ag], mol/L
The experimental data fit a straight line with a slope of -0.0660 (corresponding to Kb = 6.6·104
L/mol) and an intercept of 1.46. Thus, 1.46 = K 𝑏 ∙ N · C𝐀𝐛 = 6.6 ∙ 104 ∙ 𝑁 ∙ 1.1 ∙ 10−5 ⇒ 𝑁 = 2.
5. A clear way to determine CAb follows from the fact that the Kb value influences both the
slope and the intercept of the plot in Scatchard coordinates. As soon as Kb is determined from the
slope of the curve, CAb can be immediately calculated, provided the antibody valence N is
known. For instance, for the set A, N=1, Kb = 2·104 L/mol; CAb = 1.64 / 2·104 ≈ 82 mol/L,
which reasonably corresponds to the given value of 80 M. It can be concluded, thus, that the
ADP protein does not contain any functionally inactive antibodies or other impurities.
The same analysis for the set B is impossible, because the real enzyme valence is not known a
priori. (Value N = 2 determined above has been obtained under assumption of 100% enzyme
purity.)
1. Well known examples are: C (acethylenic carbon), S (various forms of polymeric sulfur),
Se (grey selenium), P (red phosphorus), As (black arsenic), Sb (black antimony). Not all of these
substances consist of perfectly linear chain molecules, but for sure these elements are capable of
forming quite long polymers.
2. 2HPO2− 4−
4 ⇄ P2 O7 + H2 O (ionization state of the phosphate precursor depends on pH).
3. With Pi standing for a polyphosphate with the degree of polymerization of i, for the
reaction
Pm OH + Pn OH ⇄ Pm OPn + H2 O
58
[P m +n ][H 2 O]
𝐾= .
P m OH [P n OH ]
4. The following reasons should be taken into account. First, the free energy of hydrolysis is
strongly negative, which means that the free energy of condensation (the reverse reaction) is
positive. Thus, the equilibrium constant of an elementary condensation stage is low (less than 1),
which is not consistent with the high-polymeric phosphate species. In general, lower equilibrium
concentration of (poly)phosphate molecules means that more individual condensations have
taken place, which is equivalent to the higher average degree of polymerization of the product.
This is true for process ii): lower water concentration (at a certain equilibrium constant value)
corresponds to lower equilibrium concentration of phosphate molecules (from the expression
derived in i. 3). Thus, process ii) is more favorable than i). However, process iii) is the most
favorable. According to the equation
a highly volatile HCl is formed, which is efficiently removed from the reaction mixture by
heating. As a result, the equilibrium is shifted rightwards.
Indeed, only route iii) can be applied in practice for the preparation of polyphosphoric acids.
Condensation in concentrated solutions (process ii)) is quite slow, and yields significant amounts
of polyphosphoric acids only upon heating (molten H3PO4, 230-250°C). Direct condensation in
dilute solution (process i)) is so unfavorable that may come true only when coupled with a
certain exoergic reaction (for instance, substrate phosphorylation in various biochemical
processes) with the actual mechanism much more complicated than direct condensation.
59
5.
6, 7. The Si–Cl bond is much more reactive than the C–Cl one in hydrolysis and condensation
reactions. Thus, A2 can be considered bifunctional in polycondensation reaction, giving a non-
branched polymer with the cyclic giant macromolecule of poly(chlorodimethylsiloxane) as the
final product when absolutely all Si-Cl bonds are reacted:
If hydrolysis of Si-Cl bonds is incomplete, some Cl residues are present in the polymeric
product. Incomplete condensation retains a number of OH–groups in the product.
60
Problem 28. Determination of copper and zinc by complexometric titration
2. Cu2+ ions present in the aqueous solution are reduced to Cu+ by thiosulfate. Moreover, the
latter forms with Cu+ a soluble complex [Cu(S2O3)3]5–, which is more stable than Cu2H2EDTA:
2Cu2+ + 8S2O32– = 2[Cu(S2O3)3]5– + S4O62–
3. Metal ions can be titrated with EDTA if the conditional stability constants β’ of the metal
– EDTA complexes are not less than 108–109. The β’ values are connected with the real
constants β as
β’ = αEDTA αM β,
where αEDTA and αM are molar fractions of H2EDTA2– and free metal ion, respectively. As the
values of αEDTA and αM significantly depend on pH of the solution, there is an optimal pH range
for the titration of metals. In the case of Cu2+ and Zn2+, the pH value within 5 to 6 is optimal. In
such slightly acidic medium both metals do not form hydroxy complexes (αM is high), whilst
H2EDTA2– is not further protonated (αEDTA is high).
5. The first titration (B) gives the volume of titrant VCu+Zn, whilst the second one (C) gives
VZn. Zn2+ concentration is calculated as follows:
c(Zn2+) = (VZn mL/1000 mL L-1) . cEDTA mol L-1 . 100 mL/10.00 mL . 65.39 g mol-1 / 0.1000 L
c(Zn2+) g L-1 = VZn mL . cEDTA mol L-1 . 65.39 g mol-1 . 0.1 mL-1
c(Cu2+) = ((VCu+Zn - VZn ) mL/1000 mL L-1) . cEDTA mol L-1 . 100 mL/10.00 mL . 63.55 g mol-1 /
0.1000 L
c(Cu2+) g L-1 = (VCu+Zn - VZn) mL . cEDTA mol L-1 . 63.55 g mol-1 . 0.1 mL-1
The mass ratio of the metals in alloy is calculated from c(Cu2+) and c(Zn2+) values in g L-1:
m(Cu)/m(Zn) = c(Cu2+)/c(Zn2+)
61
Problem 29. Conductometric determination of ammonium nitrate and nitric
acid
Conductivity, mS/cm
A B C
E
Conductivity, mS/cm
C
Titration of solutions of HNO3 and
NH4NO3 diluted with deionized water
(C), distilled water (D), and deionized
0 1 2 3 4 5 6 7 water containing NaCl (E).
Volume of NaOH added, mL
The difference between cases C, D, and E is due to various levels of conductivity caused by the
salts that are not titrated with NaOH.
4. Calculations can be done in the same way as for a regular acid-base titration, using titrant
volumes in inflection points VNaOH(1), VNaOH(2):
cH+Vsample = cNaOHVNaOH(1), cNH4+Vsample = cNaOH·(VNaOH(2)– VNaOH(1))
Examples.
63
A and B: if 2.45 mL of 0.9987 M NaOH spent until the inflection point, then cNH4+ is: 0.9987
2,45 = cNH4+ 25; cNH4+ = 0.0979 M.
C - E: if 2.40 mL of 0.9987 M NaOH spent until the first inflection point (neutralization of
HNO3) in a 25.0 mL sample aliquot, then cHNO3 = 0.0895 M. If the second inflection point
reached at 4.85 mL, then cNH4+ is: 0.9987 (4.85 – 2.40) = cNH4+ 25; cNH4+ = 0.0979 M.
Conductivity, mS/cm
linear decrease of conductivity due to lowering
concentration of highly mobile hydroxyl ions. After
the first equivalence point, conductivity starts
increasing due to the formation of a well
dissociating salt (a strong electrolyte) as a result of
0 1 2 3 4 5
Volume of HCl added,mL
ammonia (a weak electrolyte) neutralization. After
the second equivalence point, conductivity of the solution sharply increases due to the excess of
hydrogen ions.
1. Titration curves for a polyprotic acid (such as phosphoric acid) or a mixture of acids are
characterized by more than one endpoint if Ka1:Ka2 ≥ 104 and the equilibrium constant of acidity
of the weak acid is more than n×10–9. The equilibrium constants of acidity of phosphoric acid
are: Ka1 = 7.1×10–3, Ka2 = 6.2×10–8, Ka3 = 5.0×10–13. Thus, there are two breaks on the titration
curve of phosphoric acid (Fig. 1). The third break is not observed due to very low value of Ka3.
2. The first and second equivalence points of H3PO4 are observed at pH of about 4.7 and
9.6, respectively. For determination of hydrochloric and phosphoric acids in their mixture, one
can use indicators with color change around these pH values (for example, bromocresol green
and thymol phthalein for the first and second titrations, respectively).
4. A typical analysis of potentiometric titration data is shown in Fig. 2. The most steeply
rising portion on the curve (a) corresponds to the endpoint, which can be found more precisely
by studying dependences of the first (maximum on curve (b)) or second (zero value on curve (c))
derivatives. In the presence of ammonium salts, the reaction corresponding to the second end
point in H3PO4 titration
H2PO4– + OH– = HPO42– + H2O
is overlaid by the process
NH4+ + OH– = NH3 + H2O,
which makes the potential rise gradually rather than sharply (ammonium buffer).
65
(b) Calculation of the total amount of diammonium hydrophosphate and ammonium chloride
With VNaOH,2 designating the volume of sodium hydroxide used in titration B (that is, spent for
the neutralization of hexamethylene tetrammonium cation (CH2)6(NH+)4 obtained from the
ammonium salts), one gets:
nNH4Cl + 2nPO4 = cNaOH × VNaOH,2
The amount of phosphate n(NH4)2H3PO4 was determined in experiment A, which allows calculating
the amount of NH4Cl
nNH4Cl = cNaOH × VNaOH,2 – 2·(cHCl × VHCl – cNaOH × VNaOH,1)
and its content in the mixture:
ωNH4Cl = 10 × nNH4Cl × MNH4Cl / mmixture
66
1. Hemiaminal.
O
OH
R R R R
H N H
NH3 R'
NH2
R' H+ R'
H H
OH O
R R H+ R R
N N
H+ H2O
H R' H R'
hemiaminal
R R R R
H+ N
N
H R' R'
2. Both reactions proceed through the positively charged intermediates, iminium and
oxonium ions, respectively. While the former just loses the proton to form the final product, the
later acts as an electrophile adding another molecule of alcohol to become the full acetal.
3.
pyruvic acid
H3C CO2H H3C CO2H
H3C CO2H
N N
O
R OH R OH
H2N - H2O
N CH3 N CH3
R OH
H H
N CH3
pyridoxamine phosphate,
O
R = CH2OPO3H alanine
R OH
H3C CO2H
NH2 N CH3
pyridoxal phosphate,
R = CH2OPO3H
68
4.
O
H+
H
N
Me Me
Me Me
N
H B CN
Me Me
N
5.
H+ H2C O
H2C O
H
H
N
H H+
N N
H+ O
O
H H
H H H
H
N H
N CH2
O H 2O
H
H H
69
H
H2C O
H+
H2C O
NHMe
CH3
H3C
NHMe
H H H
O O
H
H+
H 2O H3C CH3
H3C CH3
1.
H
O N Ph
N H
N Ph
H OH
N
HO H
HO H
3 PhNhNH2 + NH3 + PhNH2
H OH
H OH
H OH
H OH
CH2OH
CH2OH
N H
N Ph
HO H
H OH
H OH
CH2OH
It is one and the same product for all the starting substances. These means the stereochemistry of
C3, C4 and C5 of the starting sugars is the same. The initial difference in nature and/or
stereochemistry at 1st and 2nd carbon atoms of the monosaccharides is equalizes by hydrazone
formation.
Ninhydrine test. The spot with the product will show no color change, whilst that with the
starting amino acid will become colored (blue-violet to brown-violet).
1.
H
OH H3C O O CH3
CH3
O
O OH
OH H3C OH CH3 OH CH3
O CH3 O
O
CH3 CH3
OH2
CH3 CH3
OH CH3 O
O
Transformation of hemiketal into full ketal needs the acid catalysis to protonate hydroxyl group,
which is further removed in the form of water molecule. The resulting positively charged
carbocation-type intermediate is stabilized by electron donation from oxygen lone pair.
71
2.
trans cis
O O
CH3 CH3
CH3 CH3
O O
cis-Fused six- and five-membered rings in the resulting product of cis-cyclohexane-1,2-diol are
more stable than trans-fused rings. The reason is the higher bond and angles distortion in trans-
fused bicycles.
3. In the furanose form of D-mannose, there is a possibility to form two rather than one (in
the pyranose from) 1,3-dioxolane rings, which is more thermodynamically favorable. Pyranose –
furanose transformation proceeds via the open aldehyde form of the carbohydrate.
5.
O
H 2N
O CH C OH
CH3 H2N CH3
CH C OH CH2
O O S
CH2
CH3 CH3 H
HS
O
O
H2N
H2N
CH C OH
CH C OH
CH3
CH3
CH2
CH2
HO S
H2O S
CH3
CH3
O
H
H3C N
CH C OH
CH2
H3C S
Acid catalysis enhances the electrophilicity of carbonyl carbon atom (enhancing carbonyl
activity). Thiol group reacts first due to higher nucleophilicity compared to that of amino group.
72
6.
O O
OH
HR R
H 2N C COOH N
OH
O
O O H O
O O
- CO2 R N
OH R
O
O O
H2O
HO
NH2
HO
-RCHO
O O
O O
N
colored product
O O
1. The viscosity values calculated from the flow times of polystyrene solutions (2 to 10 g/L)
determined with the Ubbelohde viscometer at 25 °C are given in the hereunder tables. Each flow
time value is an average of three measurements. Note that your experimental values may
significantly differ from those in the tables, since the flow times depend on the molecular
properties (mainly molecular weight distribution) of a particular polystyrene sample.
2,3,4. The intrinsic viscosity [η] can be found by either graphical extrapolation to 0
concentration (as the Y-intercept), or by linear fitting (as an absolute term) of the reduced
viscosity data.
Polystyrene solutions
0,25
0,2
Toluene
y = 0,0113x + 0,084
0,15
L/gl/g
sp/c,Vsp/c,
0,1
Methylethylketone
y = 0,0008x + 0,0313
0,05
0
0 2 4 6 8 10 12
c,c, g/l
g/L
Analysis of the data given in i. 1) leads to [η] equal to 0.0840 and 0.0313 L/g for the toluene and
methyl ethyl ketone solutions, respectively. (Three significant digits are left in both cases based
on the typical amplitude of the measured flow times).
1. Experimental flow times and the calculated specific viscosities are given in the hereunder
table.
74
Note 1. The molecular weights of the repeating units of PMMA and PEG are of 86.06 and 44.05
g/mol, respectively. Mixing of equal volumes of a 2 g/L PMMA and a 1 g/L PEG (of any
molecular weight) solutions results in a reaction mixture with the molar ratio of the PMMA and
PEG units of approximately 1:1.
Note 2. The final concentration of PMMA in the resulting mixtures and its aqueous solutions is
of 1 g/L.
2.
0.4
25 0C
40 0C
0.3
0.2
sp
0.1
0.0
0 1000 2000 3000 4000 5000 6000
PEG molecular weight
3. The reaction scheme of the complex formation is given below. A decrease of the specific
viscosity of the PMАA solution upon addition of the equimolar amount of PEG is observed,
which reflects that that polymer coils in the interpolymer complex are more compact than those
in the initial solution. The compaction is due to hydrophobization of the PMAA chain with PEG.
75
CH3 CH3 CH3 CH3
+PEO
- water release
O O
O O
Dramatic changes in the density of the complexes are observed within a rather narrow range of
PEG molecular weights (of about 1500 g/mol at 40°C and 2500 g/mol at 25 °C). Such processes
are often referred to as cooperative.
The enthalpy change in PMAA-PEG complex formation being negligible, the entropy gain due
to the release of water molecules is the driving force of the reaction.
As positions of the repeating units in a polymer chains are constrained, the total entropy of the
polymer coil is less than that of the same number of unbound monomer units. For longer
polymer chains such entropy loss is more significant. Consequently, the entropy gain as a result
of PMAA-PEG complex formation (S = S(complex) + S(water) – S(PMAA) – S(PEG)) is
increasing with an increase of the PEG chain length (total entropies of released water molecules,
the complex, and the initial PMAA molecules are nearly the same). This is why the PMAA-PEG
interaction proceeds efficiently only starting with a certain molecular weight of PEG (<1000
g/mol at 40°C and of about 1000-2000 g/mol at 25 °C).
Higher efficiency of the complex formation at elevated temperatures (PEG with a lower
molecular weight is needed to provide for a noticeable viscosity drop) contributes to the
consideration that the entropy gain is behind the process.