Modern Plastic Surgery, 2013, 3, 142-149
http://dx.doi.org/10.4236/mps.2013.34029 Published Online October 2013 (http://www.scirp.org/journal/mps)
The Combination of Parkland Formula, Using Normal
Saline, with Muir & Barclay Formula for Fluid
Resuscitation in the Initial Burn Shock Period
Medhat Emil Habib*, Said Al-Busaidi, Gihan Adly Latif, Ali Saleem Mehdi, C. Thomas
Departments of Plastic Surgery and Laboratory, Khoula Hospital, Muscat, Sultanate of Oman.
Email: *medhatemil1@hotmail.com
Received August 24th, 2013; revised September 22nd, 2013; accepted September 30th, 2013
Copyright © 2013 Medhat Emil Habib et al. This is an open access article distributed under the Creative Commons Attribution Li-
cense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
ABSTRACT
Objectives: Evaluation of the effects of withholding plasma during the initial part of the burn shock period (the shock
period in the study is estimated as the first 36 hours followi ng the burns) when it will be lost into the interstitial tissues
through the permeable capillaries. During that time crystalloids are administered. Another objective is to evaluate the
effect of administering normal saline as the crystalloid resuscitation fluid during the in itial part of the shock period. De-
sign: A Retrospective 4 years stud y compares the use of nor mal saline as the resuscitative intrav enous fluid during the
first 12 hours post burns followed by intravenous 5% Purified Plasma Protein Fraction (PPPF) during the rest of the
shock period i.e. the remaining 24 hours, with the use of the PPPF throughout the burns shock period according to Muir
and Barclay formula. Setting: The Plastic Surgery Department and the Department of Laboratory, Directorate General
of Khoula Tertiary Hospital, Muscat, Sultanate of Oman. Patients and Methods: The study included 2 groups of pa-
tients; Group A: Patients who received 5% Plasma (Human PPPF) throughout the shock period and Group B: Patients
who received crystalloids in the form of normal saline during the first 12 hours post burn followed by plasma for the
next 24 hours. Monitoring of the patients in both groups was done by using clinical signs of pulse, blood pressure, tem-
perature and urine output and by using laboratory investigations in the form of the haematocrit value, sodium, potas-
sium, chloride, total proteins and albumin levels in the blood at the time of admission and at the end of the shock period.
Results: 140 patients were included in the study; 64 in Group A and 76 in Group B. There was no mortality and the
vital signs were maintained during the shock period in both groups. The mean values of urine output were nearer to the
normal level in Group B compared to Group A. The same was observed regarding the Haematocrit value. In both
groups the mean values showed no hypoproteinaemia or hypoalbuminaemia at the end of the shock period. There was
no hypernatraemia in spite of giving 150 mmol/L of Na during the initial 12 hours post burns in Group B. The mean
values of potassium and chloride levels were normal in both groups at the end of the shock period. Conclusion: Giving
plasma during the first 12 hours of the burn shock period when the capillary leakage is maximum has no significant
benefit. The plasma usage can be reduced by 50% compared to the use of the Muir and Barclay Formula from the be-
ginning of the shock period with reduction of the costs and the possibility of transmission of undetected pathogens by
nearly the same value if crystalloids are given during the first 12 hours of burns shock period. The use of isotonic nor-
mal saline during the first 12 hours appears more appropriate as it maintains adequate sodium balance to correct the
hyponatraemia and at the same time prevents elevation of the serum potassium during the period when potassium is
released from the cells. In addition, it does not have a significant red uction on the level of the serum proteins.
Keywords: Normal Saline; Intravenous Resuscitation; Shock Period; Burn
1. Introduction
Different approaches to resuscitation of burnt patients
have been used in many Burns centers. Opinions differ
over whether to give colloids or crystalloids, what con-
centration of salt is appropriate, and so forth. With the
recent understanding of the burns pathophysiological
changes, it is logic to give the resuscitation fluids ac-
cording to these changes. During the early post-burn pe-
riod, an increased rate of capillary leakage meant that
colloids were no more effective than crystalloids and
*Corresponding a uthor.
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period 143
may in fact be potentially harmful. There is also a
breakdown of the cell membrane by the thermal injury
which injures the sodium potassium pump. This results in
increase of the intracellular sodium with hyponatraemia
and efflux of potassium. We report our experience with
the use of sodium chloride during the first 12 hours of the
burn shock period, when albumin leakage through the
capillaries is maximum, followed by plasma after resto-
ration of the capillary wall integrity.
2. Patients and Methods
This 4 years retrospective study included the patients
admitted with moderate and severe burns to the Burns
Unit of the Directorate General of Khoula Tertiary Hos-
pital, Sultanate of Oman. 140 patients were included in
the study who fulfilled the following criteria:
1) Patients admitted directly with fresh burns to the
burns unit and not referred from other hospitals later on.
2) Children with burns of >9% and adults of >15% of
their total body surface area assessed according to the
Lund and Browder chart.
The 140 patients included in the 4 years study were
divided into two groups according to the period they
were admitted to the Burns Unit. For one and half years
of the four years all the patients admitted received
Plasma only throughout the shock period and for the
other two and half years all the patients admitted during
that time received normal saline followed by Plasma.
Accordingly two group s of patients were formed:
1) Group A: Patients who received 5% Plasma (Hu-
man PPF) throughout the 36 hours of the shock period
(64 patie nts).
2) Group B: Patients who received crystalloids in the
form of normal saline during the first 12 hours post burn
followed by plasma for the next 24 hours (76 patients).
In Group A (The Plasma Group), Plasma was given
according to Muir and Barclay formula:

1ml% o f burns weightKg
2XX .
This amount was given during each of the periods of 4,
4, 4, 6, 6 and 12 hours post burn.
In Group B patients (Normal Saline and Plasma
Group), normal saline was administered during the first
12 hours post burns according to Parkland formula:
4 ml X% of burns X weight (Kg).
Half of the calculated amount was given during the
first 8 hours post burns and 1/8th was given during the
following 4 hours. (The first 4 hours of the remainin g 16
hours of the Parkland formula when the second half of
the calculated dose is supposed to be given.)
Plasma was started 12 hours post burn according to
Muir and Barclay formula:

1ml% o f burns weightKg
2XX .
This amount was given during each of the following 6
hours, 6 hours and 12 hours.
Daily requirement fluids were added to both groups of
patients.
Monitoring of the patients was done using the clinical
signs and laboratory investigations. The clinical signs
included, the pulse, blood pressure, temperature and
urine output. The laboratory investigations included the
haematocrit value, sodium, potassium, chloride, total
proteins and albumin levels in the blood at the time of
admission and at the end of the shock period manage-
ment.
3. Results
140 patients were included in this study.
87 patients sustained scald burns (62%), 46 patients
sustained flame burns (33%), 4 patients sustained electric
burns (3%) while 3 patients sustained chemical burns
(2%) as shown in Figure 1.
The percentage of the burnt area to the total body sur-
face area ranged between 9% - 90% with a mean value of
24.9%.
The number of children and adults and also the gender
in each group are shown in Table 1.
The mean age in Group A was 17.8 years (Range from
9 months to 57 years) while in Group B it was 15.05
years (Range from 8 months to 61 years).
There was no mortal ity during the s hock period in bot h
groups.
The clinical evaluation showed that the vital signs
were maintained in both groups. There was no significant
difference in mean values of the heart rate in both groups.
In children, it was 121/min. in Group A and 119.3/min.
in Group B while in adults it was 88.3/min. in Group A
and 89.5/min. in Group B.
62%
33%
2% 3%
Scald
Fl a me
Chem i ca l
Electric
Figure 1. The percentage of the types of burns.
Table 1. The distribution of children and adults in each
group and the gender distri bution.
Child Adult Male Female
Group A Plasma 37 27 41 23
Group B N. S. & Plasma48 28 49 27
Total 85 55 90 50
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period
144
The mean values of the blood pressure in both groups
showed also no significant difference. In children, it was
107/52 mmHg in Group A and 109/57 mmHg in Group
B while in adults it was 130/65 mmHg in Group A and
125/62 mmHg in Group B.
The mean values of the temperature were elevated in
both groups. While in Group A it was 39.4 C it was 38.2
C in Group B (Table 2).
The urine output was elev ated in both the groups but it
was nearer to the normal side in Group B (3 ml/kg/hr for
children and 1.9 ml/kg/hr for adults) compared to Group
A (3.2 ml/kg/hr for children and 4 ml/kg/hr for adults)
(Figure 2).
The haematocrit values (normally mean values 38%
for children, 45% for adult males and 41% for adult fe-
males) were measured at the time of admission before
starting the shock period management and at the end of
the shock period 36 hours post burns. The results of the
mean haematocrit values measured at the end of the
shock period in children and adults in both groups
showed that it was nearer to normal in Group B (34.6 %
in children and 43.2% in adults) compared to the results
of Group A (32.3% in children and 36.2% in adults)
(Table 3).
In Group A, 9.4% of the patients (6 patients) were
found having hypoalbuminaemia (Normal value 35 - 53
gm/l) at the time of admission. At the end of the shock
period, this percentage increased to 15.6% (10 patients).
The levels were higher in Group B; 13.2% (10 patients)
at the time of admission and 31.6% (24 patients) after the
Table 2. The mean values of the vital signs in each group.
Heart rate/min. Blood pressure mmHg
Children Adults Children Adults
Temp.
C
Group
A 121 88.3 107/52 130/65 39.4
Group
B 119.3 89.5 109/57 125/62 38.2
0
0.5
1
1.5
2
2.5
3
3.5
4
Child Adult
Pl a sma
N.S.&Plasma
Figure 2. The mean values of the urine output of children
and adults in both the groups.
shock period.
No patients were admitted with initial hyperalbu-
minaemia in both groups while one patient in each group
had postinfusion hyperalbuminaemia (1.6% in Group A
and 1.3% in Group B) (Figure 3).
The mean values of albumin were maintained within
normal limits in both groups (40.6 gm/l at the time of
admission and 40.3 gm/l at the end of the shock period in
Group A and 39.3 gm/l at th e time of admission and 37.1
gm/l at the end of the shock period in Group B) (Table
4).
The percentage of abnormal total protein levels was
higher than those of the abnormal albumin levels (Nor-
mal value 60 - 83 gm/l). In Group A, 10.9% of the pa-
tients (7 patients) were found having hypoproteinaemia
at the time of admission. At the end of the shock period,
this percentage increased to 42.2% (27 patients). The
levels were near in Group B; 7.9% (6 patients) at the
time of admission and 47.4% (36 patients) after the
shock period.
One patient was admitted with initial hyperproteinae-
mia in Group B. No patients were admitted with initial
hyperproteinaemia in Group A and no patients developed
postinfusion hyperproteinaemia in both the groups (Fig-
ure 4).
The mean values of total protein were also maintained
within normal limits in bo th groups (70.1 gm/l at the time
of admission and 62.5 gm/l at the end of the shock period
Table 3. The mean haematocrit values at the time of admis-
sion and after the shock periods (S. P.) in adults and chil-
dren in both groups.
Before
S. P. After
S. P. Before
S. P. After
S. P.
Group
Child Adult
Group A Plasma 36.4 32.3 44.5 36.2
Group B N. S. & Plasma38 34.6 43.8 43.2
0%
5%
10%
15%
20%
25%
30%
35%
PlasmaN.S.&
Plasma
Initial hypoalbuminaemia
Postin fu s i on h ypoalbumi n aemia
Initial hyperalbuminaemia
Postinfusion hyperalbuminaemi a
Figure 3. Percentage of abnormal albumin levels in both
groups.
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period 145
Table 4. Mean values of albumin and total protein levels in
both groups at the time of admission and at the end of the
shock period.
Albumin
(35 - 53 gm/l) Total protein
(60 - 83 gm/l)
Group
Before S. P. After S. P. Before S. P. After S. P.
Group A
Plasma 40.6 40.3 70.1 62.5
Group B
N. S. &
Plasma 39.3 37.1 69.6 60
0%
5%
10%
15%
20%
25%
30%
35%
40%
45%
50%
Plasma N.S.&
Plasm
a
Intitial hypoproteinaemia
Post inf usio n hy poprot ei na em ia
In itial hype rproteinaemia
Post infusio n h y perprotein a emia
Figure 4. Percentage of abnormal protein levels in both
groups.
in Group A and 69.6 gm/l at the time of admission and
60 gm/l at the end of the shock period in Group B) (Ta-
ble 4).
The percentage of patients who were having hypona-
traemia at the time of admission and at the end of the
shock period were high in both groups (Normal values
137 - 148 mmol/L). The number of patients admitted
with hyponatraemia in Group A at the time of admission
was 36 patients (56.3%) and the number increased to 38
patients at the end of the shock period (59.4%). In Group
B, the patients who were having hyponatraemia at the
time of admission were 40 patients (52.6%). The number
increased to 45 patients at the end of the shock period
(59.2%). No patients were admitted with initial hyperna-
traemia in Group A but one developed postinfusion hy-
pernatraemia at the end of the shock period (1.6%). No
patients were having hypernatraemia at the time of ad-
mission or at the end of the shock period in Group B
(Figure 5).
The mean values show return of the sodium level to
normal at the end of shock period in Group A (increased
from 136.4 mmol/L to 137.1 mmol/L after resuscitation)
while in Group B it remained below the normal values
and even reduced from 136.8 mmol/L to 136.1 mmol/L
(Table 5).
In Group A, 23 patients (35.9%) were admitted with
0%
10%
20%
30%
40%
50%
60%
PlasmaN .S. &
Plasm
a
Initial hyponatraemia
Postinfussi on h yponatraem i a
Initial hype rnatraemia
Postinfusi on h ypernatraem i a
Figure 5. The percentage of patients with abnormal Sodium
levels at the time of admission and at the end of the shock
period in both groups.
Table 5. The mean values of Sodium, Potassium and Chlo-
ride levels before and after resuscitation in both groups.
Na (137 -
148 mmol/l) K (3.6 -
5 mmol/l) Cl (101 -
111 mmol/l)
Group Pre
Res. Post
Res. Pre
Res. Post
Res. Pre
Res. Post
Res.
Plasma 136.4137.1 3.7 3.8 106 107.3
N. S. & Plasma136.8136.1 3.6 3.7 105.7106.6
initial hypokalaemia and the situation improved after
resuscitation to 18 patients (28.1%) with hypokalaemia.
This percentage was better in Group B (28 patients ad-
mitted with hypokalaemia 36.8% and 18 patients had
hypokalaemia at the end of the shock period management
23.9%). While in Group A one patient was admitted with
initial hyperkalaemia (1.6%) and no patients had post
infusion hyperkalaemia, the opposite was there in Group
B as no patients were admitted with initial hyperk alaemia
but one patient developed it post infusion (1.3%) (Figure
6).
The mean values showed normal potassium levels in
both groups before and after the infusion (3.7 mmol/l
increased to 3.8 mmol/l in Group A and 3.6 mmol/l in-
creased top 3.7 mmol/l in Group B) (Table 5).
Regarding the chloride level in the blood, 2 patients in
Group A were admitted with initial hypochloraemia
(3.1%) and this level increased to 4 patients after resus-
citation (6.25%). In Group B, 7 patients were admitted
with initial hypochloraemia (9.2%) and the situation im-
proved by the end of resuscitation to 4 patients (5.3%).
The patients admitted with initial hyperchloraemia in
Group A were 7 patients (10.9%) and increased to 13
patients at the end of the shock period (20.3%). Similarly,
the patients who had increased chloride level in Group B
increased from 6 patients (7.9%) at the time of admission
to 13 patients (17.1%) after the resuscitation period (Fig-
ure 7).
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period
146
0%
5%
10%
15%
20%
25%
30%
35%
40%
PlasmaN.S.&
Plasma
Intial hypokalaemia
Postinfusion hypokalaemia
Intial h yperkalaem i a
Postinfusion hyperkalaemia
Figure 6. The percentage of patients with abnormal Potas-
sium levels at the time of admission and at the end of the
shock period in both groups.
0%
5%
10%
15%
20%
25%
Plasma N.S.&
Plasma
Initial hypochloraemia
Postinfusion hypochloraemia
Initial hype rchloraemia
Postinfusion hyperchloraemia
Figure 7. The percentage of patients with abnormal Chlo-
ride levels at the time of admission and at the end of the
shock period in both groups.
The mean values showed maintenance of the chloride
levels within normal values in both groups before and
after resuscitation (106 mmol/l in Group A increased to
107.3 mmol/l and 105.7 mmol/l in Group B increased to
106.6 mmol/l) (Table 5).
4. Discussion
The development of effective fluid resuscitation regi-
mens is one of the cornerstones of modern burn treatment
and perhaps the advance which has most directly im-
proved patient survival. Some of the early used regimens
were introduced by Evans [1], Moyer [2], Arturson [3],
Monafo [4], Shires [5], Baxter [6] and others.
The understanding of the pathophysiological changes
which occur in burns was progressing through the years.
Studies showed an increase in the microvascular perme-
ability which results in leakage of fluid, electrolytes and
proteins from the intravascular space into the interstitial
space, impairing tissue perfusion. The loss of proteins
into the interstitial space rapidly decreases the intravas-
cular colloid osmotic pressure, and results in a reversed
osmotic gradient [7]. This disturbs the natural Starling
forces and leads to oedema formation [8]. Lund et al.
claimed that a negative interstitial pressure develops in
the thermally injured skin. This pressure constitutes a
strong “suction” adding markedly to the oedema gener-
ating effect of increased capillary permeability [9].
There is a considerable debate about the type of fluid
to be used for resuscitation of burn patients [10,11].
Those who prefer colloid resuscitation cite the advan-
tages of prolonged intravascular half-life, better intra-
vascular volume effect and maintenance of plasma col-
loid osmotic pressure with reduction of complications of
fluid overlo ad [12-15]. Those who opp ose the use of col-
loids warn about their extravascular distribution via in-
creased microvascular permeability and their side effects
on coagulation and kidney function [16,17]. Risk of 4% -
6% mortality was seen when colloids were used in criti-
cally ill patients [18,19]. The opponents for the use of
crystalloids argue that crystallo ids can cause a significan t
rise in total fluid volume required with increased tissue
and pulmonary oedema. Those who support crystalloid
use claim that although there is a risk of increased tissue
and pulmonary edema, no clinically significant sequelae
have been demonstrated and the cost-benefit of crystal-
loids outweighs theoretical advantages of colloids [20].
Considering the recent understanding of the patho-
physiological changes which occur in burns, there was an
increased trend to withhold colloids until capillary per-
meability reverts to normal, using crystalloid only till
that stage [21]. Cocks et al. mentioned that many burn
protocols restrict the use of colloids in the resuscitation
of burns shock patients before 12 hours post burn to
avoid the leakage of colloids from the capillaries with
their harmful effects. During the initial period of burns
crystalloids are given [22].
In our study we used in Group B patients only crystal-
loids during the fi rst 12 hours of the burn shock pe riod at
the time when the capillary leakage is high and cannot
withhold the colloids. After 12 hours collo ids in the form
of 5% Albumin were started for the following 24 hours.
We administered crystalloids during the first 12 hours
of burns according to the Parkland Formula as it is the
most commonly used formula nowadays even in UK and
Ireland [23]. We calculated the amount of crystalloids in
the first 24 hours of burns and half the calculated amount
was given in the initial 8 hours followed by 1/8th of the
calculated amount in the following 4 hours. This amount
completed the first 12 hours according to Parkland for-
mula. The following 24 hours of burn were managed
according to the Muir and Barclay formula after skipping
the first 12 hours of the original formula i.e. colloids
were given according to the percentage of burns multi-
plied by the body weight and divided by 2 for the fol-
lowing 6 hours, 6 hours and 12 hours.
The original Parkland formula introduced by Baxter
included administration of 5% albumin for 8 hours after
the first 24 hours of the burns [6,24,25]. Modern itera-
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period 147
tions of the Parkland formula have omitted the colloid
bolus. This resulted in more fluid requirement than is
predicted by the formula and the appearance of the phe-
nomena of “fluid creep”. Use of colloids as a routine
component of resuscitation of burns patients following
the crystalloids showed to reduce this phenomena and
reduce the complications of overloading the circulation
with crystalloids to compensate for the fluid deficit and
reduced the possibility of abdominal compartment syn-
drome [26,27]. Chung et al., during the Operation Iraqi
Freedom developed a protocol to utilize 5% albumin
solution in any patient whose 24-fluid requirements are
projected to exceed 6 ml/kg/% TBSA. This has reduced
the incidence of compartment syndrome to zero [28].
Yowler and Fratianne resuscitated burned patients with
albumin at 12 hours post-burn when the fluid require-
ments were found greater than 150% of that predicted by
formula. They found that this can reduce the total fluid
requirements and the burn edema [29]. In our study we
started the colloids after 12 hours period of crystalloid
infusion (Group B patients). The mean values of the
pulse, blood pressure and urine output did not show
manifestations of fluid creep. The mean value of urine
output specifically was markedly above the normal range
which meant adequate hydration of the patients with no
manifestations of oliguria requiring additional fluid ad-
ministration.
The intravenous fluid used in Parkland Formula is
Ringer’s lactate. Ringer lactate is the commonest fluid
used on a world-wide basis [30].
Normal saline was used in Evan’s formula. Intrave-
nous input of saline was begun by Reiss in 1880 [31].
Dulhunty et al., in their resuscitation of the 80 patients
included in their study used Hartmann’s solution (58/80;
72.5%) and normal saline (21/80; 26.3%) [32].
In our study we preferred to use normal saline and not
Ringer Lactate solution in Group B patients for the fol-
lowing reasons:
1) Normal saline contains 150 mmol/L sodium while
Ringer Lactate contains only 131 mmol/L of the sodium
(the normal range of sodium in plasma is 137 - 148
mmol/l). In a situation when there is influx of sodium
inside the cells with resulting hyponatraemia, it is better
to supplement the deficient sodiu m with 150 mmol/L and
not a fluid with a lower sodium level than the plasma
level.
2) Ringer Lactate contains 5 mmol/L of potassium
ions. The normal plasma level of potassium is 3.6 - 5
mmol/L. In a situation in which there is release of potas-
sium ions from the cells with possibility of resulting hy-
perkalaemia, it is preferred to give a fluid clear from po-
tassium during that time than to add more potassium to a
possibl e h y p erkalaemic sit ua tion.
3) Ringer lactate contains 29 mmol/L bicarbonate as
lactate. Bicarbonate administration was found recently to
be hazards especially if the burnt patient is developing
acidosis. Bicarbonate shifts the oxyhaemoglobin disso-
ciation curve to the left inhibiting the release of oxygen
to the tissues. It may produce carbon dioxide which may
cross into the cell causing intracellular acidosis (para-
doxic acidosis) [33]. Another drawback in administering
ringer lactate solution is related to glucose metabolism.
Intracellular glucose is metabolized to pyrovate through
the Embden-Meyerhof pathway of glucose metabolism.
If oxygen is available, the pyrovate enters the mitochon-
dria and is further metabolized by oxidative phosphory-
lation to carbon dioxid e and hydrogen ato ms through the
Krebs cycle releasing energy. If oxygen is not available,
pyrovate accumulates outside the mitochondria and is
converted to lactate. Accumulation of lactate can cause
acidosis [34]. On their study on 166 burned patients,
Kamolz et al., found that a better chance of survival oc-
curs when resuscitation results in normal values of lac-
tate than supra-normal values [35].
When we compared giving colloids throughout the
first 36 hours (shock period) following the burns (Group
A patients) with giving crystalloids for 12 hours followed
by colloids (Group B patients) we found nearly no dif-
ference in the mean values of the vital signs of both
groups. There was no mortality during the burn shock
period in both the groups.
The urine output was elevated in both groups but it
was nearer to the normal side in Group B (3 ml/kg/hr
for children and 1.9 ml/kg/hr for adults) compared to
Group A (3.2 ml/kg/hr for children and 4 ml/kg/hr for
adults. Normal values 0.5 - 1 ml/kg/hr in adults and 1 - 2
ml/kg/hr in children) [36,37].
Comparing also the haematocrit value in both groups,
the Group B patients were nearer to normal after the I. V.
fluid replacement in adults and children compared to
Group A patie nt s indi cat i ng adeq uat e hy d rati on.
The number of patients who developed postinfusion
hypoalbuminaemia was higher in Group B compared to
Group A patients but the mean values of albumin were
maintained within normal limits in both groups. One pa-
tient in Group B developed postinfusion hyperalbu-
minaemia.
The percentage of patients who developed post infu-
sion hypoprotinaemia was slightly higher in Group B
compared to Group A patients but the mean values of
total protein were maintained within normal limits in
both groups (62.5 gm/l at the end of the shock period in
Group A and 60 gm/l at the end of the shock period in
Group B). The above data showed that administering
crystalloids followed by colloids did not result in hypo-
albuminaemia or hypoprotinaemia according to the mean
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period
148
values.
In spite of administering 150 mmol/L of sodium in the
sodium chloride solution for the initial 1 2 hours of resus-
citation (Group B patients), no patients developed post-
infusion hypernatraemia. At the same time the mean
values of sodium postinfusion in Group B were even
below the normal level compared to Group A patients
which were within the normal range. These data show
that administering sodium chloride does not cause hy-
pernatraemia.
In spite of administering no potassium ions during the
first 12 hours post burn in Group B patients, the percent-
age of patients who developed postinfusion hypokalae-
mia in this group was less than those in Group A. One
patient in Group B developed post infusion hyperkalae-
mia. The mean values in both groups show normal levels
of potassium at the end of the shock period in both
groups.
In spite of administering 150 mmol/L of chloride in
the sodium chloride solution for the initial 12 hours of
resuscitation (Group B patients), the percentage of pa-
tients who developed postinfusion hyperchloraemia was
less than those in Group A. The mean values showed
maintenance of the chloride levels within normal range
in both groups before and after resuscitation.
5. Conclusions
Using Normal Saline in the initial part of the burn shock
period is consistent with the understanding of the patho-
physiology of burns and the changes that occur during its
phases.
Withholding plasma during the first 12 hours when the
capillary leakage is maximum will reduce the loss of
albumin into the interstitial space.
The use of isotonic normal saline during the first 12
hours appears more appropriate as it maintains adequate
sodium balance and prevents elevation of the serum po-
tassium. In addition, there was no significan t reduction in
the level of serum proteins.
REFERENCES
[1] E. Evans, O. Purnell, P. Robinett, A. Batchelor and M.
Martin, “Fluid and Electrolyte Requirements in Severe
Burns,” Annals of Surgery, Vol. 135, No. 6, 1952, p. 804.
http://dx.doi.org/10.1097/00000658-195206000-00006
[2] C. A. Moyer, H. W. Margraf and W. W. Monafo, “Burn
Shock and Extravascular Sodium Deficiency: Treatment
with Ringer’s Solution with Lactate,” Archives of Surgery,
Vol. 90, 1965, pp. 799-811.
http://dx.doi.org/10.1001/archsurg.1965.01320120001001
[3] G. Arturson, “Pathophysiological Aspects of the Burn
Syndrome with Special Reference to Liver Injury and Al-
terations of Capillary Permeability,” Acta Chirurgica
Scandinavica. Supplementum, Vol. 274, 1961, pp. 1-135.
[4] W. W. Monafo, “The Treatment of Burn Shock by the
Intravenous and Oral Administration of Hypertonic Lac-
tated Saline Solution,” Journal of Trauma, Vol. 10, No. 7,
1970, pp. 575-586.
http://dx.doi.org/10.1097/00005373-197007000-00006
[5] G. T. Shires, C. J. Carrico, C. R. Baxter, A. H. Giesecke
and M. T. Jenkins, “Principles in Treatment of Severely
Injured Patients,” Advances in Surgery, Vol. 4, 1970, pp.
255-324.
[6] C. R. Baxter and T. Shires, “Physiological Response to
Crystalloid Resuscitation of Severe Burns,” Annals of the
New York Academy of Sciences, Vol. 150, No. 3, 1968, pp.
874-894.
http://dx.doi.org/10.1111/j.1749-6632.1968.tb14738.x
[7] S. Tricklebank, “Modern Trends in Fluid Therapy for
Burns,” Burns, Vol. 35, No. 6, 2009, pp. 757-767.
http://dx.doi.org/10.1016/j.burns.2008.09.007
[8] Guyton and Hall, “Textbook of Medical Physiology,” 9th
Edition, 1996, pp. 187-193.
[9] T. Lund, H. Onarheim and R. K. Reed, “Pathogenisis of
Edema Formation in Burn Injuries,” World Journal of
Surgery, Vol. 16, No. 1, 1992, pp. 2-9.
http://dx.doi.org/10.1007/BF02067107
[10] C. Holm, “Resuscitation in Shock Associated with Burns.
Tradition or Evidence-Based Medicine?” Resuscitation,
Vol. 44, No. 3, 2000, pp. 157-164.
http://dx.doi.org/10.1016/S0300-9572(00)00159-3
[11] A. R. Webb, “Crystalloid or Colloid for Resuscitation.
Are We Any the Wiser?” Critical Care, Vol. 3, No. 3,
1999, pp. 25-28. http://dx.doi.org/10.1186/cc346
[12] L. Fodor, A. Fodor, Y. Ramon, O. Shoshani, Y. Rissin
and Y. Ullmann, “Controversies in Fluid Resuscitation
for Burn Management: Literature Review and Our Ex-
perience,” Injury, Vol. 37, No. 5, 2006, pp. 374-379.
http://dx.doi.org/10.1016/j.injury.2005.06.037
[13] M. J. Davies, “Crystalloid or Colloid: Does It Matter?”
Journal of Clinical Anesthesia, Vol. 1, No. 6, 1989, pp.
464-471.
http://dx.doi.org/10.1016/0952-8180(89)90013-5
[14] M. I. Griffel and B. S. Kaufman, “Pharmacology of Col-
loids and Crystalloids,” Critical Care Clinics, Vol. 8, No.
2, 1992, pp. 235-253.
[15] A. Aharoni, D. Abramovici, M. Weinberger, R. Moscona
and B. Hirshowitz, “Burn Resuscitation with a Low-Vol-
ume Plasma Regimen—Analysis of Mortality,” Baillières
Clinical Anaesthesiology, Vol. 11, 1997, pp. 369-384.
[16] R. S. Bisonni, D. R. Holtgrave, F. Lawler and D. S. Mar-
ley, “Colloids versus Crystalloids in Fluid Resuscitation:
An Analysis of Randomized Controlled Trials,” Journal
of Family Practice, Vol. 32, No. 4, 1991, pp. 387-390.
[17] D. C. Gore, J. M. Dalton and T. W. Gehr, “Colloid Infu-
sions Reduce Glomerular Filtration in Resuscitated Burn
Victims,” Journal of Trauma, Vol. 40, No. 3, 1996, pp.
356-360.
http://dx.doi.org/10.1097/00005373-199603000-00005
[18] G. Schierhout and I. Roberts, “Fluid Resuscitation with
Copyright © 2013 SciRes. MPS
The Combination of Parkland Formula, Using Normal Saline, with Muir & Barclay formula
for Fluid Resuscitation in the Initial Burn Shock Period
Copyright © 2013 SciRes. MPS
149
Colloid or Crystalloid Solutions in Critically Ill Patients:
A Systematic Review of Randomized Trials,” BMJ, Vol.
316, No. 28, 1998, pp. 961-964.
http://dx.doi.org/10.1136/bmj.316.7136.961
[19] I. Roberts, “Cochrane Injuries Group Albumin Reviewers.
Human Albumin Administration in Critically Ill Patients:
Systematic Review of Randomised Controlled Trials,”
BMJ, Vol. 317, No. 7153, 1998, pp. 235-340.
http://dx.doi.org/10.1136/bmj.317.7153.235
[20] G. V. Sudhakar, and P. Lakshmi, “Role of HES 130/ 0.4
in Resuscitation of Patients with Major Burn Injury,”
Transfusion Alternatives in Transfusion Medicine, Vol.
10, No. 2, 2008, pp. 43-50.
http://dx.doi.org/10.1111/j.1778-428X.2008.00107.x
[21] R. Cole, “The UK Albumin Debate,” Burns, Vol. 25, No.
7, 1999, p. 565.
[22] A. J. Cocks, A. O’Connell and H. Martin, “Crystalloids,
Colloids and the Kids: A Review of Paediatric Burns in
Intensive Care,” Burns, Vol. 24, No. 8, 1998, pp. 717-724.
http://dx.doi.org/10.1016/S0305-4179(98)00102-8
[23] S. Tricklebank, “Modern Trends in Fluid Therapy for
Burns,” Burns, Vol. 35, No. 6, 2009, pp. 757-767.
http://dx.doi.org/10.1016/j.burns.2008.09.007
[24] M. Haberal, E. S. Abali and H. Karakayali, “Fluid Man-
agement in Major Burn Injuries,” Indian Journal of Plas-
tic Surgery, Vol. 43, Suppl. 1, 2010, pp. 29-36.
http://dx.doi.org/10.4103/0970-0358.70715
[25] G. D. Warden, “Burn Shock Resuscitation,” World Jour-
nal of Surgery, Vol. 16, No. 1, 1992, pp. 16-23.
http://dx.doi.org/10.1007/BF02067109
[26] C. Csontos, V. Foldi, T. Fischer and L. Bogar, “Factors
Affecting Fluid Requirement on the First Day after Se-
vere Burn Trauma,” ANZ Journal of Surgery, Vol. 77, No.
9, 2007, pp. 745-748.
http://dx.doi.org/10.1111/j.1445-2197.2007.04221.x
[27] J. R. Saffle, “The Phenomenon of ‘Fluid Creep’ in Acute
Burn Resuscitation,” Journal of Burn Care & Research,
Vol. 28, No. 3, 2007, pp. 382-395.
[28] K. K. Chung, L. H. Blackbourne, S. E. Wolf, et al.,
“Evolution of Burn Resuscitation in Operation Iraqi Free-
dom,” Journal of Burn Care & Research, Vol. 27, No. 5,
2006, pp. 606-611.
http://dx.doi.org/10.1097/01.BCR.0000235466.57137.f2
[29] C. J. Yowler and R. B. Fratianne, “Current Status of Burn
Resuscitation,” Clinics in Plastic Surgery, Vol. 27, No. 1,
2000, pp. 1-10.
[30] A. D. Mansfield, “Resuscitation and Monitoring,” Resus-
citation, Vol. 44, No. 3, 2000, pp. 157-164.
[31] D. Mahler, A. Baruchin, D. Hauben, A. R. Moscona, B.
Hirshowitz, H. Y. Kaplan, I. Peled, M. R. Wexler, E.
Vure and J. Shulman, “Recent Concepts Regarding the
Resuscitation of the Burned Patient,” Burns, Vol. 9, No. 1,
1982, pp. 30-37.
http://dx.doi.org/10.1016/0305-4179(82)90133-4
[32] J. M. Dulhunty, R. J. Boots, M. J. Rudd, M. J. Muller and
J. Lipman, “Increased Fluid Resuscitation Can Lead to
Adverse Outcomes in Major-Burn Injured Patients, but
Low Mortality Is Achievable,” Burns, Vol. 34, No. 8,
2008, pp. 1090-1097.
http://dx.doi.org/10.1016/j.burns.2008.01.011
[33] R. Touquet, J. Fothergill and M. W. Platt, “Resuscita-
tion,” In: R. M. Kirk, A. O. Mansfield and J. P. S. Coch-
rane, Eds., Clinical Surgery in General, 3rd Edition, 1999,
pp. 3-18.
[34] W. F. Ganong, “Review of Medical Physiology; Section
IV Endocrinology, Metabolism and Reproductive Func-
tion; 17 Energy Balance,” Metabolism and Nutrition,
1995, pp. 255-289.
[35] L. P. Kamolz, H. Andel, W. Schramm, G. Meissl, D. N.
Herndon and M. Frey, “Lactate: Early Predictor of Mor-
bidity and Mortality in Patients with Severe Burns,”
Burns, Vol. 31, No. 8, 2005, pp. 986-990.
http://dx.doi.org/10.1016/j.burns.2005.06.019
[36] J. K. Rose and D. N. Herndon, “Advances in the Treat-
ment of Burn Patients,” Burns, Vol. 23, Suppl. 1, 1997,
pp. S19-S26.
[37] G. D. Warden, “Burn Shock Resuscitation,” World Jour-
nal of Surgery, Vol. 16, No. 1, 1992, pp. 16-23.
http://dx.doi.org/10.1007/BF02067109