Accepted Manuscript Guiding principles of fluid and volume therapy Dita Aditianingsih, M.D. Yohanes W.H. George, M.D.

PII:

S1521-6896(14)00050-0

DOI:

10.1016/j.bpa.2014.07.002

Reference:

YBEAN 813

To appear in:

Best Practice & Research Clinical Anaesthesiology

Received Date: 4 May 2014 Revised Date:

20 June 2014

Accepted Date: 4 July 2014

Please cite this article as: Aditianingsih D, George YWH, Guiding principles of fluid and volume therapy, Best Practice & Research Clinical Anaesthesiology (2014), doi: 10.1016/j.bpa.2014.07.002. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Guiding principles of fluid and volume therapy

RI PT

Dita Aditianingsih, M.D.*, Yohanes W.H. George, M.D.

Department of Anaesthesia and Intensive Care, University of Indonesia, Cipto

SC

Mangunkusumo Hospital, Jakarta, Indonesia

*Corresponding author. Department of Anaesthesia and Intensive Care, University of

City, 10430, Indonesia.

M AN U

Indonesia, Cipto Mangunkusumo Hospital, Diponegoro St. No. 71, Central Jakarta

Tel: 62-21-3143736/ 62-815-1819244 Fax: 62-21-3912526

TE D

E-mail address: [email protected] (D. Aditianingsih).

AC C

EP

[email protected] (Y. George)

1

ACCEPTED MANUSCRIPT

Abstract

RI PT

Fluid therapy is a core concept in the management of peri-operative and critically ill patients for maintenance of intravascular volume and organ perfusion. Recent evidence regarding the vascular barrier and its role in terms of vascular

SC

leakage has led to a new concept for fluid administration. The choice of fluid used should be based on the fluid composition and the underlying pathophysiology of the

M AN U

patient. Avoidance of both hypo- and hypervolaemia is essential when treating circulatory failure. In daily practice, the assessment of individual thresholds in order to optimize cardiac preload and avoid hypovolaemia or deleterious fluid overload remains a challenge. Liberal vs restrictive fluid management has been challenged by

TE D

recent evidence, and the ideal approach appears to be goal-directed fluid therapy.

Keywords:

EP

‘double-barrier concept’

AC C

crystalloids colloids

liberal vs restrictive fluid management preload responsiveness goal-directed fluid therapy

2

ACCEPTED MANUSCRIPT

Introduction

RI PT

Fluid therapy is a core concept in the management of peri-operative and critically ill patients. Several aspects influence the benefits and side effects of fluid administration. Recent evidence regarding the vascular barrier, its physiological

SC

relevance and its role in terms of vascular leakage has led to a new concept for intravascular fluid administration. In contrast to Starling’s principle of fluid exchange,

function of the vascular barrier.

M AN U

it has been suggested that endothelial glycocalyx may play a pivotal role in the

Traditionally, high volumes of crystalloids have been administered to patients undergoing major surgery, based on presumptions of pre-operative dehydration and

TE D

overestimation of intra-operative fluid loss due to the ‘third space’. The Surviving Sepsis Campaign recommends goal-directed fluid therapy (GDT) in the case of persistent hypoperfusion within 6 h. Inadequate fluid replacement in cases of

EP

hypovolaemia results in reduced cardiac output (CO) and decreased oxygen delivery

AC C

(DO2) to tissues, leading to organ dysfunction. However, excessive fluid administration and a positive fluid balance are associated with various complications and increased risk of mortality. Studies have compared restrictive fluid administration with liberal fluid administration for volume management, and tried to find an association between the physiological function of the vascular barrier and the type of fluid administered (i.e. crystalloid or colloid). However, the results of these studies are conflicting. Recent studies have focused on the use of advanced haemodynamic

3

ACCEPTED MANUSCRIPT monitoring, and proposed the rational concept of GDT to improve outcomes in critically ill patients and those undergoing high-risk surgery. This brief review will summarize the relevant physiology of body fluid distribution, the role of the endothelial vascular barrier, and the effects of various

RI PT

intravenous solutions. Therapeutic goals will be highlighted, along with critical

for clinicians.

Physiology and pathophysiological basics

SC

evaluation of haemodynamic parameters to guide fluid therapy, and recommendations

M AN U

The human body consists of 60% water; approximately two-thirds is located in the intracellular compartment and one-third is located in the extracellular compartment. The extracellular compartment consists of the interstitial space, blood plasma and small amounts of secreted transcellular fluid (e.g. intra-ocular,

TE D

cerebrospinal, gastrointestinal). The capillary endothelium is freely permeable to water, electrolytes, nutriments and glucose, but is impermeable to large molecules such as proteins and colloids, which are confined to the intravascular space.

EP

Intravenous fluid therapy targets the intravascular fluid volume (IVFV, blood

AC C

volume), the extracellular fluid volume (ECFV, extracellular space) or both. The composition and differential use of intravenous fluids should be determined by the fluid space targeted, and there appears to be no difference between the intra-operative, peri-operative, postoperative and intensive care settings. Volume replacement aims to replace a reduction in IVFV and to correct hypovolaemia in order to maintain haemodynamics and vital signs. This is achieved with an essentially physiological solution that contains colloid osmotic components, that is both iso-oncotic and isotonic. Fluid replacement aims to offset any impending or existing ECFV deficit

4

ACCEPTED MANUSCRIPT due to loss of cutaneous, enteral or renal fluid. This is achieved with a physiological solution that contains all osmotically active components and is isotonic. Electrolyte replacement or osmotherapy aims to restore the physiological total body fluid volume [intracellular fluid volume (ICFV) plus ECFC] when loss of cutaneous, enteral or

RI PT

renal fluid has altered the composition and/or volume of either or both fluid spaces (ICFV and/or ECFV).1

Fluid moves between the intravascular compartment and the extravascular

SC

compartment across the vascular endothelial barrier, and movement is classified as physiological and pathological. Physiological fluid movement occurs continuously

M AN U

through an intact vascular endothelial barrier. Fluid is redistributed slowly between the interstitial and intracellular compartments, and returns to the intravascular compartment via the lymphatic system. Physiological fluid movement does not cause interstitial oedema but fluid loss can result in dehydration. Pathological fluid

TE D

movement occurs when the vascular endothelial barrier is damaged. This type of movement allows fluid accumulation leading to interstitial oedema, and can lead to acute hypovolaemia if the fluid loss is excessive.2

EP

Ernest Starling described fluid movement as a net balance of colloid osmotic

AC C

pressure and hydrostatic pressure between the intravascular and interstitial compartments. The fluid components of the intravascular compartment are mainly pulled inwards due to the colloid osmotic pressure produced by the protein content of plasma. This pressure opposes the high intravascular hydrostatic pressure, which has a tendency to push fluid out of the vessels into the interstitium. This principle suggests that both the interstitial colloid osmotic pressure and hydrostatic pressure are far below the intravascular pressure, and the net result of these forces is a small outward leak of fluid and proteins from the vasculature into the interstitium; this is returned to

5

ACCEPTED MANUSCRIPT the blood vessels continually via the lymphatic system.3 This classic principle suggests that the endothelial cell layer alone is responsible for the function of the vascular barrier.4 In contrast to Starling’s principle, it has been suggested that endothelial glycocalyx may act as a primary molecular filter, generating an effective

RI PT

oncotic gradient at the endothelial microstructural level, with the intravascular– interstitial protein gradient concentration playing a minor role.5,6 This new ‘doublebarrier concept’ suggests that the ‘first-line’ endothelial glycocalyx layer and

SC

endothelial cells comprising the endothelial surface layer maintain the vascular barrier. Together, these layers have a thickness of 0.4–1.2 µm, and function in

M AN U

dynamic equilibrium with approximately 800–1000 ml of circulating and noncirculating plasma in humans. A normal level of plasma albumin is required for optimal function. 7-9 The endothelial surface layer/glycocalyx layer is the first contact surface between blood and tissue. It addition to its role as a vascular barrier, it is

TE D

involved in processes such as inflammation, haemostasis, coagulation and regulation of vasomotor tone. If damaged, glycocalyx loses much of its ability to act as a barrier, and causes platelet aggregation, leukocyte adhesion and increased transendothelial

AC C

EP

permeability, resulting in interstitial oedema.10

Fluids for volume therapy: crystalloids or colloids Crystalloids

Crystalloids are distributed freely across the vascular endothelial barrier. A

paradigm exists that four times as much crystalloid compared with colloid is needed for the same volume effect, and several studies have shown comparable results. Isotonic or near-isotonic crystalloids, such as normal saline (NS), lactated Ringer’s (LR) solution and acetated Ringer’s solution, have shown that the distribution phase

6

ACCEPTED MANUSCRIPT takes 25–35 min. The increase in plasma volume during infusion is greater than commonly suggested. Following infusion of 2 l of acetated Ringer’s solution over 30 min, 50% was located in the plasma at the end of infusion in normovolaemic volunteers; other studies have reported an increase in plasma volume of 65–70% after

RI PT

infusion of 1.1 l of acetated Ringer’s solution over 10 min and 2 l of acetated Ringer’s solution over 20 min.11-13 In patients undergoing general anaesthesia, up to 60% of acetated Ringer’s solution was located in the plasma during continuous infusion.14 In

SC

healthy female volunteers, the fraction of infused Ringer’s solution that remains in the plasma is higher for slower infusion rates and decreases with infusion time. The

M AN U

increase in plasma volume after 30 min of infusion is 50–75%. A relatively long period of time is required for crystalloid fluids to distribute; as such, slow infusions are more effective than bolus administrations.15

The effects of different types of crystalloids have been studied in healthy

TE D

volunteers. LR solution decreased serum osmolality briefly, with a return to baseline after 1 h. NS did not affect serum osmolality but caused metabolic acidosis.16 Serum albumin was decreased by dilution but returned to baseline, indicating redistribution

EP

within fluid compartments. The decrease in albumin level lasted for more than 6 h

AC C

with NS, but returned to normal within 1 h with dextrose 5%. Haemoglobin was also decreased by dilution; the water content from the dextrose 5% infusion was excreted after 2 h, but NS had a longer-lasting effect with only 30% of sodium and water excreted after 6 h.17 The increase in plasma volume, as estimated by dilution of haemoglobin and albumin, was shown to be more sustained with NS (56% of infused volume at 6 h) compared with Hartmann’s solution (30%). There were no significant differences in serum potassium, sodium, urea or total osmolality, but NS resulted in a decrease in bicarbonate and sustained hyperchloraemia for more than 6 h.18

7

ACCEPTED MANUSCRIPT The use of large volumes of NS causes hyperchloraemic metabolic acidosis. Twenty-nine percent of patients with shock developed hyperchloraemic acidosis within 24 h of infusion of more than 1 l of NS in 1 h or less. There is a strong likelihood of hyperchloraemic acidosis if patients receive a high volume (at least 4 l)

RI PT

of NS at a high infusion rate.19 In short-term interventions, limiting NS administration with the use of colloids in any carrier has a moderate hyperchloraemic effect, and this is relatively transient for patients with normal organ function. Healthy human

SC

volunteers need more than 2 days to excrete a salt and water load of 2 l of NS.16,18,20,21 Worse outcomes were observed in patients with acute illness or critical illness,

M AN U

where the ability to excrete a salt and water load is impaired. A significant decrease in the incidence of acute kidney injury (AKI) and the need for renal replacement therapy (RRT) were observed when hyperchloraemic solutions including NS were avoided.22,23 The mechanism of renal toxicity caused by NS or other hyperchloraemic

TE D

solutions remains unclear, but a high chloride load could reduce renal function by brief provocation of the tubule–glomerular system. NS worsened sepsis-induced AKI due to increased inflammation, and this outcome was confirmed by decreased

AC C

EP

function, increased levels of injury biomarkers and histological results.24

Colloids

Colloids are fluids that contain macromolecules exceeding 40 kDa, and are

classified as natural (e.g. albumin) and artificial (e.g. starches, dextrans, gelatins). Colloids are dissolved in either saline or more balanced salt solutions. Colloid gelatins of low-to-medium molecular weight and albumin are more likely to leak into the interstitial compartment compared with hydroxyethyl starch (HES), which has a

8

ACCEPTED MANUSCRIPT higher molecular weight. HES is retained intravascularly for longer, and the duration of retention in the circulatory system differs between colloids.25 Colloids are believed to increase plasma oncotic pressure. Due to their higher molecular weight, colloids may interact and adsorb to the glycocalyx layer, and

RI PT

restrict ultrafiltration, while crystalloids equilibrate rapidly between the intravascular and interstitial compartments.26 In healthy, intact vascular barriers, colloids remain in the intravascular compartment for up to 16 h, compared with 30–60 min for

SC

crystalloids such as LR solution and NS. The colloid volume required to achieve a haemodynamic target is less compared with the crystalloid volume (1:4), and

M AN U

therefore intravascular volume remains increased for longer.27 Endothelial glycocalyx is often damaged in injured and critically ill patients, and therefore capillary leakage is more likely. When a colloid solution diffuses into the interstitium, it reduces the oncotic pressure gradient between the capillary barriers, and results in further fluid

TE D

extravasation. At best, colloids can restore intravascular volume within minutes or hours, but their effect on total body water is cumulative and persists. Critically ill

fluid.28,29

EP

patients tend to retain more fluid, and need days or weeks to excrete the excess

AC C

Albumin is the main molecule used to preserve intravascular osmotic pressure, and is an ideal colloid to restore protein deficits in the vasculature. Human albumin 4– 5% in saline is a natural colloid for volume resuscitation. It is produced by the fractionation of blood, and is heat-treated to prevent transmission of pathogens; as such, it may cause allergic and immunological complications. The Saline vs Albumin Fluid Evaluation (SAFE) study showed no significant benefits in terms of mortality rate at 28 days or development of new organ failure. Additional analysis in predefined subgroups showed that, in patients with traumatic brain injury, there was a significant

9

ACCEPTED MANUSCRIPT association between albumin administration and higher mortality rate at 2 years; and, in patients with severe sepsis, there was an association between decreased mortality rate at 28 days and albumin administration, suggesting a potential benefit of albumin resuscitation.30-32 In 2013, a Cochrane review found no evidence that colloid

RI PT

resuscitation (including albumin) reduces the risk of morbidity or mortality compared with crystalloid resuscitation in heterogenous critically ill patients.33

Gelatins are polydispersed polypeptides from degraded bovine collagen. The

SC

average molecular weight of gelatins is 30–35 kDa, they have comparable volumeexpanding capability, and they are relatively safe in terms of coagulation and organ

M AN U

integrity except for kidney function. Patients undergoing aortic aneurysm surgery who were resuscitated with 4% gelatin demonstrated more distinct tubular damage with a higher level of serum urea and creatinine compared with patients resuscitated with HES, and concern regarding the risk of AKI associated with the use of gelatins was

TE D

raised.34 The use of gelatins has not been studied in a high-quality randomized controlled trial (RCT). Consequently, given the lack of clinical benefits and potential

impairment.35

EP

for nephrotoxicity, the use of gelatins should be limited in patients with renal

AC C

HES is an artificial polymer derived from hydroxyethyl substitution of amylopectin obtained from sorghum, waxy maize or potatoes. Hydroxyethylation of the glucose unit protects against hydrolysis degradation by non-specific amylases and water solubility in the blood. This semisynthetic colloid is available with various concentrations, molecular weights, molar substitutions, C2/C6 ratios, solvents and pharmacological profiles. HES with a high molecular weight (>200 kDa) and a molar substitution ratio >0.5 (200/0.6) is associated with coagulopathy due to a reduction in von Willebrand factor and factor VIII, leading to decreased platelet adhesion, changes

10

ACCEPTED MANUSCRIPT in viscoelasticity and fibrinolysis.36 HES with a higher molecular weight is metabolized more slowly and causes prolonged intravascular expansion, but may accumulate in subcutaneous tissue, liver and kidney. The use of 10% HES 200/0.5 in patients with sepsis was associated with increased incidence of AKI and need for

RI PT

RRT.37 Hyperoncotic colloids induce glomerular filtration of hyperoncotic molecules, resulting in hyperviscous urine and stasis of tubular flow, causing ‘osmotic nephrosislike lesions’ in the kidneys.38

SC

Newer HESs have a lower molecular weight (130 kDa) and molar substitution ratios of 0.38–0.45. The recommended maximum daily dose of HES 130/0.4 is 33–50

M AN U

ml.kg-1.39 HES 130/0.4 has been shown to attenuate inflammatory responses when administered as a resuscitation fluid in a septic shock and haemorrhagic shock rat model, by decreasing the levels of tumour necrosis factor-alpha, interleukins and oxidative stress.40,41 Two large RCTs have evaluated the safety of HES for patients

TE D

with severe sepsis and septic shock. The 6S trial is an RCT involving 800 patients with severe sepsis and septic shock. The investigators reported that the use of 6% HES (130/0.42), compared with acetated Ringer’s solution, was associated with

EP

increased incidence of AKI (Kidney Disease Improving Global Outcome stage 2–3 criteria), increased need for RRT, increased overall mortality at 30 days and

AC C

significantly increased mortality at 90 days.42 The CHEST trial is an RCT involving 7000 patients with severe sepsis. The investigators found that the use of 6% HES (130/0.4), compared with NS, was associated with lower incidence of AKI, and was not associated with a significant difference in mortality at 90 days. However, this trial found that use of HES was associated with increased urine output in patients with low risk for AKI, but increased levels of serum creatinine in patients with higher risk for AKI.43 Both the 6S trial and the CHEST trial found a significant increase in the need

11

ACCEPTED MANUSCRIPT for RRT associated with HES administration, and no significant difference in shortterm haemodynamic resuscitation targets between HES and crystalloids. The CHEST trial found that the use of HES was associated with the use of approximately 30% less fluid, faster elevation of central venous pressure (CVP) and lower incidence of new

RI PT

cases of shock. The HES:crystalloid ratio in the CHEST trial was 1:1.3, which is similar to the albumin:saline ratio in the SAFE study. The different results from these two trials are likely to be due to the fact that patients in the CHEST trial were in a

SC

better state to deal with hazards, while patients in the 6S trial were sicker, had been resuscitated adequately prior to study enrolment, and the 6S trial was designed to use

M AN U

the same volume of fluid and did not allow a decrease in resuscitation volume in both arms.20,30,42,43 The Pharmacovigilance Risk Assessment Committee of the European Medicines Agency concluded that, until further evidence is available, HES must no longer be used in patients with sepsis, burn injuries or critical illness because of the

TE D

risk of AKI and mortality. HES is contra-indicated for use in patients with severe coagulopathy, and patients with renal impairment or needing RRT, and use of HES must be discontinued at the first sign of coagulopathy or AKI. HES should only be

EP

used for rapid volume replacement due to acute blood loss at the lowest effective dose

AC C

for the shortest period of time, when crystalloids alone are not considered sufficient. Administration should be guided by continuous haemodynamic monitoring, and infusion should be stopped as soon as appropriate haemodynamic goals have been achieved.44

Liberal vs restrictive fluid administration Traditionally, in peri-operative settings, a large volume of crystalloid solution is infused routinely to achieve an adequate intravascular volume. This approach is

12

ACCEPTED MANUSCRIPT based on the concept that patients tend to be hypovolaemic pre-operatively due to prolonged fasting, cathartic bowel preparation, perspiration and urinary output. General anaesthesia and neuraxial blockade result in hypotension, and this often triggers liberal fluid administration to achieve the haemodynamic target. This should

RI PT

be treated by vasoactive agents.45,46

In minor surgery, peri-operative fluid shifts are small and the risk of organ dysfunction is low. Pre-operative fasting accounts for fluid deficits, and more liberal

SC

administration of crystalloids to healthy patients undergoing moderate-risk surgery led to a better recovery profile compared with patients who received restricted

M AN U

amounts of the same crystalloid.47,48 The concept of the ‘third space’ in surgical exposure leads to aggressive fluid replacement of the insensible loss. As a result, many postoperative patients have a positive fluid balance, and the risk of complications increases. No significant data exist to support the concept of the ‘third

TE D

space’, and a positive fluid balance was associated with poorer outcome.49 A review of patients undergoing major abdominal surgery and knee arthroplasty, excluding high-risk patients, compared liberal and restrictive fluid regimens; it concluded that it

EP

is difficult to define ‘liberal’ or ‘restrictive’ protocols in clinical practice, because the

AC C

studies varied in terms of design, types of fluid administered, additional fluids administered, outcome variables, definitions of intra- and postoperative periods, and the fact that a restrictive regimen in one study may be liberal in another.50,51 Liberal fluid administration in patients undergoing major surgery was associated with increased risk of pneumonia, longer time to bowel movement and increased length of hospital stay, compared with the restrictive therapy and GDT groups.52 Patients undergoing moderate-risk surgery seem to benefit from more liberal fluid

13

ACCEPTED MANUSCRIPT administration, while patients undergoing high-risk or major surgery seem to benefit from restrictive or conservative strategies. In sepsis, distributive shock and oedema are attributed to a combination of increased capillary permeability to proteins and increased net transcapillary

RI PT

hydrostatic pressure due to reduced precapillary vasoconstriction. Early GDT is a stepwise approach that improved 30-day mortality in patients with sepsis, using central venous oxygen saturation (ScvO2) >70% as an additional endpoint as well as

SC

optimized CVP and mean arterial pressure (MAP).53 The GDT arm received, on average, 2 l more fluid than the control arm in the first 6 h, although overall fluid

M AN U

volume in the first 72 h was similar in both groups. In a subset analysis, patients with sepsis who had renal support including fluid removal at enrolment had a lower mortality rate and a shorter duration of mechanical ventilation despite administration of equal volumes.53,54 The Fluids and Catheter Treatment Trial determined that

TE D

conservative fluid management 24 h after the establishment of acute respiratory distress syndrome could significantly improve lung and central nervous system function, decrease the need for sedation, reduce the duration of mechanical ventilation

EP

and reduce the length of stay in an intensive care unit (ICU). However, the study suggested that a conservative fluid management strategy should be applied cautiously

AC C

during the resuscitation phase.55 An early adequate fluid management strategy in the resuscitation phase corrects global tissue hypoxia (reflected by decreased lactate and increased ScvO2), and decreases the incidence of mechanical ventilation in the first 72 h. Fluid resuscitation improves microcirculation and modulates certain distinct biomarkers in the early phase of sepsis, but not in the late phase of sepsis.56 The Sepsis Occurrence in Acutely Ill Patients (SOAP) study and Vasopressin and Septic Shock Trial (VASST) favour a conservative fluid management strategy. The SOAP

14

ACCEPTED MANUSCRIPT study showed that a positive fluid balance within 72 h was associated with severity and mortality in the subgroup of patients with severe sepsis and septic shock.57 The VASST indicated that higher CVP at 12 h and a higher positive fluid balance on day 4 was associated with increased risk of mortality in patients with septic shock.58 Despite

RI PT

potential biases such as indications, time dependencies, repeated exposure of interventions, competing risks, and exclusion of the most sick and least sick patients, most studies found that a positive fluid balance was a risk factor in patients with

SC

sepsis. In patients with septic shock and acute respiratory distress syndrome, higher initial intravenous fluid volumes followed by a negative fluid balance for 2

M AN U

consecutive days within the first 7 days of shock resulted in a lower mortality rate.59 However, rather than relying on ‘liberal’ or ‘restrictive’ terms, it is more important to focus on the precise volume and timing of fluid administration based on targets for

TE D

correcting hypovolaemia with GDT.

Goal-directed volume therapy

An imbalance between DO2 and oxygen consumption (VO2) is common in

EP

patients undergoing high-risk surgery and critically ill patients. Tissue oxygen supply

AC C

can be determined by DO2, whereas optimization of CO is the primary factor to match metabolic demands. Fluid resuscitation has been regarded as the first step in optimizing CO, but numerous studies have reported that only 50% of haemodynamically unstable patients responded to fluid challenges.60,61 MAP, heart rate and diuresis are measured routinely but cannot assess haemodynamic instability or differentiate its causes accurately. Traditionally, for perioperative fluid administration, the predefined values of CVP and pulmonary arterial occlusion pressure (PAOP) are used to estimate left atrial pressure as an assumption

15

ACCEPTED MANUSCRIPT of left ventricular preload. Targets are achieved by fluid infusion and combinations of inotropes. Shoemaker et al. introduced the concept of targeting supranormal values of CO and DO2–VO2 using a pulmonary arterial catheter in patients undergoing highrisk surgery, and this was found to be associated with better outcomes in these

RI PT

patients.62 However, increased mortality was observed in critically ill patients treated with supranormal target values; therefore, ScvO2 and oxygen extraction ratio were added as resuscitation targets.63 Early GDT used CVP, MAP and ScvO2 as target

SC

parameters for initial fluid resuscitation in patients with severe sepsis and septic shock.53 Unfortunately, both CVP and PAOP are poor markers of intravascular

M AN U

volume, primarily due to non-linear variations in vascular compliance, and do not correlate with circulating blood volume. A systematic review verified that CVP and PAOP were poor indicators of preload and volume responsiveness to changes in stroke volume or CO.64

TE D

The first step in a fluid resuscitation strategy is the assessment of CO and preload. The Frank–Starling curve is used to determine the relationship between ventricular preload and stroke volume in individual patients, and refers to static

EP

volumetric parameters such as cardiac preload and dynamic parameters to predict

AC C

preload responsiveness of patients (Fig. 1).65,70 The parameters can be evaluated best using continuous monitoring methods, ranging from classical thermodilution techniques with a pulmonary artery catheter and transpulmonary indicator dilution, to less invasive non-calibrated pulse contour analysis of arterial pressure signal, and oesophageal Doppler. Static parameters, such as global end-diastolic volume (GEDV) and intrathoracic blood volume (ITBV), can be determined by transpulmonary thermodilution techniques. These parameters are not limited by spontaneous breathing, and have been shown to be valuable indicators of cardiac preload.66

16

ACCEPTED MANUSCRIPT Dynamic parameters, such as pulse pressure variation (PPV), stroke volume variation (SVV) and pulse variability index (PVI), are derived from interactions between the heart and the lungs during mechanical ventilation. Arterial PPV is the variation in arterial pulse pressure during positive-pressure-controlled ventilation, and SVV is the

RI PT

variation in stroke volume during positive-pressure-controlled ventilation; both are calculated using pulse contour analysis of the area beneath the arterial waveform curve or by oesophageal Doppler monitoring measurements. PPV and SVV use the

SC

magnitude of the respiratory change in arterial pulse pressure and stroke volume index as indicators of preload dependence, and are reliable predictors of preload

M AN U

responsiveness. PVI is calculated continuously by a non-invasive device that measures change in the perfusion index (ratio of non-pulsatile to pulsatile blood flow through the peripheral capillary bed) during a respiratory cycle. Variations in PPV, SVV ≥10–13% and PVI ≥15% are highly predictive of preload responsiveness. SVV,

TE D

PPV and PVI are unreliable in the presence of spontaneous breathing, cardiac arrhythmia, open chest, tidal volume 7–10 ml.kg-1) values were found in non-survivors, and this was an independent risk factor for mortality at 28 days.73,74 EVLW is a valuable tool to guide fluid management, and appears to be a good

EP

predictor in critically ill patients.

AC C

Three forms of shock (hypovolaemic, obstructive and cardiogenic) are associated with reduced CO; therefore, these conditions have a positive effect on overall outcome after normalization of CO. In contrast, in cases of distributive shock (e.g. septic shock), inadequate tissue perfusion and cellular dysfunction persist in the presence of normal or even elevated CO. This defect occurs due to microcirculatory alterations and shunting, and is characterized by persistent regional dysoxia as signified by elevated lactate levels, increased P(cv-a)CO2 (difference between arterial PaCO2 and central venous PvCO2) and high ScvO2. Lactate is the product of anaerobic

18

ACCEPTED MANUSCRIPT metabolism when tissues are being hypoperfused.53,60 Several studies have shown a significantly lower lactate level in the GDT arm postoperatively.75 GDT guided by close lactate monitoring showed that a fluid-restricted regime in patients undergoing major elective gastrointestinal surgery may lead to fluid insufficiency and low tissue

RI PT

perfusion in up to 28% of patients. Close monitoring of serum lactate levels and lactate clearance to guide fluid administration may improve early detection of hypoperfusion and evaluation of the patient’s response to fluid therapy.53,76 P(cv-

SC

a)CO2 is a blood-flow-related blood gas parameter. Arterial PaCO2 is dependent on pulmonary gas exchange, but central venous PvCO2 is dependent on the ability of the

M AN U

flow or CO to wash out CO2 from the tissue. Therefore, increased P(cv-a)CO2 may help the early detection of hypoperfusion and can guide fluid administration.77,78 In patients with sepsis, microcirculatory impairment and tissue hypoxia persist despite systemic haemodynamic optimization, such as normal or increased CO,

TE D

normal blood flow and normal DO2. Microcirculatory alterations are a significant risk factor in this population. A study in patients with septic shock found reduced microcirculatory perfusion, and improvement of the microcirculatory flow response to

EP

fluids might be a good target for fluid therapy. Targeting the microcirculation in fluid

AC C

resuscitation with direct bedside observations using near-infrared spectroscopy, orthogonal polarization spectral and sidestream darkfield imaging can evaluate and guide GDT at microcirculatory level, and result in more optimal resuscitation.79

Conclusions The choice and timing of administration of different types of fluid, and the amount, should be based on the fluid composition and the underlying pathophysiology of the patient. Early adequate and late conservative fluid

19

ACCEPTED MANUSCRIPT management strategies use macro- and microcirculatory parameters as the targets of resuscitation, and aim to match the increased oxygen demands during stress and systemic inflammatory responses, and avoid further complications. Therapeutic interventions should be made early to restore fluid balance, preserve tissue perfusion

RI PT

and tissue oxygenation. It is achieved by optimizing macrocirculatory haemodynamic targets such as static pressure targets (e.g. MAP, CVP, PAOP), volumetric targets (CO, SV, GEDV, ITBV), dynamic parameter targets (e.g. PPV, SVV, PVI), organ-

SC

function target (e.g. EVLW), in parallel with improvement of microcirculatory targets (lactate, ScvO2, P(cv-a)CO2) and microcirculatory flow. When performed early during

M AN U

or after surgery, or after recognition of shock, in appropriate patients, and using welldefined targets, GDT has been shown to improve the outcomes of patients undergoing high-risk surgery and critically ill patients.

-

TE D

Practice points

The new ‘double-barrier concept’ and fluid kinetics of endothelial glycocalyx suggest that glycocalyx degradation may lead to increased capillary

High chloride solutions should be avoided in high volume resuscitation due to

AC C

-

EP

permeability and interstitial oedema.

hyperchloraemic metabolic acidosis and nephrotoxicity. It is advisable to use balanced electrolyte solutions for volume replacement.

-

HES solutions must be avoided in critically ill patients at high risk of AKI and coagulopathy, and should only be used for the treatment of hypovolaemia due to acute blood loss, at the lowest effective dose for the shortest period of time, when crystalloids alone are not considered sufficient.

-

Both hypo- and hypervolaemia have an adverse effect on patient outcome.

20

ACCEPTED MANUSCRIPT GDT based on the patient’s underlying pathophysiology, guided by volumetric parameters (e.g. GEDV, ITBV), dynamic parameters (e.g. PPV, SVV, PVI, PLR test, EEO test), and organ function target (e.g. EVLW), represents the ideal macrocirculatory approach. Bedside

observations

using

near-infrared

spectroscopy,

orthogonal

RI PT

-

polarization spectral and sidestream darkfield imaging can guide GDT at

SC

microcirculatory level and result in more optimal resuscitation.

Research agenda

Future high-quality trials are needed to determine whether colloids,

M AN U

-

crystalloids or a combination of the two fluid types could produce better outcomes, especially in particular clinical conditions. -

Further prospective RCTs are needed to evaluate whether a late conservative

TE D

strategy can improve outcomes, and the optimum time for implementation, especially in specific populations of critically ill patients. -

Large multicentre trials are needed to compare various peri-operative GDT

EP

protocols in terms of peri-operative morbidity and mortality among patients in

AC C

all risk strata.

References

1. Kaye A, Riopelle J. Intravascular fluid and electrolyte physiology. Miller R, Miller’s Anesthesia, vol. 54, 7th Edition. Philadelphia: Churchill Livingstone 2009; 1705–1707. 2. Chappell D, Jacob M, Hofmann-Kiefer K, et al. A rational approach to perioperative fluid management. Anesthesiology 2008; 109: 723–740.

21

ACCEPTED MANUSCRIPT 3. Strunden MS, Heckel K, Goetz AE, Reuter D. Perioperative fluid and volume management: physiological basis, tolls and strategies. Annal of Intensive Care 2011; 1: 1-9. 4. Starling E. On the absorption of fluid the connective tissue spaces. Journal of

RI PT

Physiology 1896; 19: 312-326.

5. Adamson RH, Lenz JF, Zhang X, et al. Oncotic pressures opposing filtration across non-fenestrated rat microvessels. Journal of Physiology 2004; 557:

SC

889-907.

6. Weinbaum S. Whitaker distinguished lecture: models to solve mysteries in

M AN U

biomechanics at the cellular level; a new view of fiber matrix layers. Annals of Biomedical Engineering 1998; 26: 627–643.

7. Jacob M, Chappell D, Rehm M. The ‘third space’—fact or fiction? Best Practice & Research Clinical Anaesthesiology 2009; 23: 145–157.

TE D

8. Rehm M, Zahler S, Lotsch M, et al. Endothelial glycocalyx as an additional barrier determining extravasation of 6% hydroxyethyl starch or 5% albumin solutions in the coronary vascular bed. Anesthesiology 2004; 100: 1211–1223.

EP

9. Jacob M, Bruegger D, Rehm M, et al. The endothelial glycocalyx affords

AC C

compatibility of Starling’s principle and high cardiac interstitial albumin levels. Cardiovascular Research 2007; 73: 575–586.

10. Ait-Oufella H, Maury E, Lehoux S, et al. The endothelium: physiological functions and role in microcirculatory failure during severe sepsis. Intensive

Care Medicine 2010; 36: 1286–1298. 11. Drobin D, Hahn RG. Volume kinetics of Ringer’s solution in hypovolemic volunteers. Anesthesiology 1999; 90: 81–91. 12. Svense´n CH, Rodhe PM, Olsson J, et al. Arteriovenous differences in plasma

22

ACCEPTED MANUSCRIPT dilution and the distribution kinetics of lactated Ringer’s solution. Anesthesia and Analgesia 2009; 108: 128–133. 13. Norberg Å, Hahn RG, Husong Li, et al. Population volume kinetics predicts

volunteers. Anesthesiology 2007; 107: 24–32.

RI PT

retention of 0.9% saline infused in awake and isoflurane-anesthetized

14. Hahn RG. Volume effect of Ringer solution in the blood during general anaesthesia. European Journal of Anaesthesiology 1998; 15: 427–432.

SC

15. Hahn RG, Drobin D, Ståhle L. Volume kinetics of Ringer’s solution in female volunteers. British Journal of Anaesthesia 1997; 78: 144–148.

M AN U

16. Williams EL, Hiltebrand KL, McCormick SA, Bedel MJ. The effect of intravenous lactated Ringer’s solution versus 0.9% sodium chloride solution on serum osmolality in human volunteers. Anesthesia and Analgesia 1999; 88: 999–1003.

TE D

17. Lobo DN, Stanga Z, Simpson JA, et al. Dilution and redistribution effects of rapid 2-litre infusions of 0.9% (w/v) saline and 5% (w/v) dextrose on haematological parameters and serum biochemistry in normal subjects: a

EP

double-blind crossover study. Clinical Science 2001; 101: 173–179.

AC C

18. Reid F, Lobo DN, Williams RN, et al. (Ab)normal saline and physiological Hartmann’s solution: a randomized double-blind crossover study. Clinical

Science 2003; 104: 17–24.

19. Gheorghe C, Dadu R, Blot C, et al. Hyperchloremic metabolic acidosis following resuscitation of shock. Chest 2010; 138: 1521–1522. 20. Myburgh JA, Mythen MG. Resuscitation fluids. The New England Journal of Medicine 2013; 369: 1243-1251. 21. O’Malley CM, Frumento RJ, Hardy MA, et al. A randomized, double-blind

23

ACCEPTED MANUSCRIPT comparison of lactated Ringer’s solution and 0.9% NaCl during renal transplantation. Anesthesia and Analgesia 2005; 100: 1518–1524. 22. Shaw AD, Bagshaw SM, Goldstein SL, et al. Major complications, mortality, and resource utilization after open abdominal surgery: 0.9% saline compared

RI PT

to Plasma-Lyte. Annals of Surgery 2012; 255: 821–829.

23. Yunos NM, Bellomo R, Hegarty C, et al. Association between a chlorideliberal vs chloride-restrictive intravenous fluid administration strategy and

Association 2012; 308: 1566–1572.

SC

kidney injury in critically ill adults. The Journal of the American Medical

M AN U

24. Zhou F, Peng Z, Kellum JA. Effects of resuscitation fluids on experimental sepsis-induced acute kidney injury. Critical Care Medicine 2011; 38: Supplement: A13.

25. Traylor RJ, Pearl RG. Crystalloid versus colloid versus colloid: all colloids are

TE D

not created equal. Anesthesia and Analgesia 1996; 83: 209–212. 26. Levick JR, Michel CC. Microvascular fluid exchange and the revised Starling principle. Cardiovascular Research 2010; 87: 198–210.

AC C

38.

EP

27. O’Neill D. The right plasma volume expander. Nursing Times 2001; 97:27–

28. Prowle JR, Bellomo R. Fluid administration and the kidney. Current Opinion in Critical Care 2010; 16: 332–336.

29. Shaw AD, Kellum JA. The risk of AKI in patients treated with intravenous solutions containing hydroxyethyl starch. Clinical Journal of the American Society of Nephrology 2013; 8: 497–503.

30. The SAFE Study Investigators. A comparison of albumin and saline for fluid resuscitation in the intensive care unit. The New England Journal of Medicine

24

ACCEPTED MANUSCRIPT 2004; 350: 2247-2256. 31. The SAFE Study Investigators. Saline or albumin for fluid resuscitation in patients with traumatic brain injury. The New England Journal of Medicine 2007; 357: 874-84.

RI PT

32. Finfer S, McEvoy S, Bellomo R, et al. Impact of albumin compared to saline on organ function and mortality of patients with severe sepsis. Intensive Care Medicine 2011; 37: 86-96.

SC

33. Perel P, Roberts I, Ker K. Colloids versus crystalloids for fluid resuscitation in critically ill patients. Cochrane Database of Systematic Reviews 2013; 2:

M AN U

CD000567.

34. Mahmood A, Gosling P, Vohra RK. Randomized clinical trial comparing the effects on renal function of hydroxyethyl starch or gelatin during aortic aneurysm surgery. British Journal of Surgery 2007; 94: 427–433.

TE D

35. Bayer O, Reinhart K, Sakr Y, et al. Renal effects of synthetic colloids and crystalloids in patients with severe sepsis: a prospective sequential comparison. Critical Care Medicine 2011; 39: 1335-1342.

EP

36. Franz A, Braünlich P, Gamsjäger T, et al. The effects of hydroxyethyl starches

AC C

of varying molecular weights on platelet function. Anesthesia and Analgesia 2001; 92: 1402-1407.

37. Brunkhorst FM, Englel C, Bloos F, et al. German competence network sepsis (SepNet): Intensive insulin therapy and pentastarch resuscitation in severe sepsis. The New England Journal of Medicine 2008, 358:125-39. 38. Chinitz JL, Kim KE, Onesti G, Swartz C. Pathophysiology and prevention of dextran-40- induced anuria. The Journal of Laboratory and Clinical Medicine 1971, 77: 76-87.

25

ACCEPTED MANUSCRIPT 39. McSwain NE, Champion HR, Fabian TC, et al. State of the art of fluid resuscitation 2010: prehospital and immediate transition to the hospital. The Journal of Trauma and Acute care Surgery 2011; 70: Supplement: S2-S10. 40. Feng X, Yan W, Wang Z, et al. Hydroxyethyl starch, but not modified fluid

RI PT

gelatin, affects inflammatory response in a rat model of polymicrobial sepsis with capillary leakage. Anesthesia and Analgesia 2007; 104: 624-630.

41. Wang P, Li Y, Li J. Hydroxyethyl starch 130/0.4 prevents the early pulmonary

SC

inflammatory response and oxidative stress after hemorrhagic shock and resuscitation in rats. International Immunopharmacology 2009; 9:347-353.

M AN U

42. Perner A, Haase N, Guttormsen AB, et al. Hydroxyethyl starch 130/0.42 versus Ringer’s acetate in severe sepsis. The New England Journal of Medicine 2012; 367: 124-134.

43. Myburgh JA, Finfer S, Bellomo R, et al. Hydroxyethyl starch or saline for

TE D

fluid resuscitation in intensive care. The New England Journal of Medicine 2012; 367: 1901-1911. 44.

Hydroxyethyl-starch solutions (HES) no longer to be used in patients with

AC C

2013.

EP

sespsi or burn injuries or crtically ill patients. European Medicine Agency

45. Jackson R, Reid JA, Thorburn J. Volume preloading is not essential to prevent spinal-induced hypotension at Caesarean section. British Journal of Anaesthesia 1995; 75: 262–265.

46. Jacob M, Chappell D, Conzen P,et al. Blood volume is normal after preoperative overnight fasting. Acta Anaesthesiologica Scandinavica 2008; 52: 522–529. 47. Lambert KG, Wakim JH, Lambert NE. Preoperative fluid bolus and reduction

26

ACCEPTED MANUSCRIPT of postoperative nausea and vomiting in patients undergoing laparoscopic gynecologic surgery. American Association of Nurse Anesthetist Journal 2009; 77:1 10–14. 48. Holte K, Klarskov B, Christensen DS, et al. Liberal versus restrictive fluid

RI PT

administration to improve recovery after laparoscopic cholecystectomy: a randomized, double-blind study. Annals of Surgery 2004; 240: 892–899.

49. Brandstrup B, Svensen C, Engquist A. Hemorrhage and operation cause a

SC

contraction of the extracellular space needing replacement-evidence and implications? A systematic review. Surgery 2006; 139: 419–432.

M AN U

50. Brandstrup B, Tonnesen H, Beier-Holgersen R, et al. Effects of intravenous fluid restriction on postoperative complications: comparison of two perioperative fluid regimens: a randomized assessor-blinded multicenter trial. Annals of Surgery 2003; 238: 641–648.

TE D

51. Bundgaard-Nielsen M, Secher NH, Kehlet H. ‘Liberal’ vs. ‘restrictive’ perioperative fluid therapy—a critical assessment of the evidence. Acta Anaesthesiologica Scandinavica 2009; 53: 843–851.

EP

52. Corcoran T, Rhodes JE, Clarke S, Myles P et al. Perioperative Fluid

AC C

Management Strategies in Major Surgery: A Stratified Meta-Analysis. Anesthesia and Analgesia 2012; 114:640-651.

53. Rivers E, Nguyen B, Havstad S, et al. Early goal-directed therapy in the treatment of severe sepsis and septic shock. The New England Journal of Medicine 2001; 345: 1368–1377. 54. Donnino M, Shirazi E, Wira C, et al. Early goal-directed therapy in patients with end-stage renal disease. Critical Care 2004; 8:163.

55. Wiedemann HP, Wheeler AP, Bernard GR, et al. Comparison of two fluid

27

ACCEPTED MANUSCRIPT management strategies in acute lung injury. The New England Journal of Medicine 2006; 354: 2564–2575. 56. Rivers EP, Kruse JA, Jacobsen G, et al. The influence of early hemodynamic

Care Medicine 2007; 35: 2016–2024.

RI PT

optimization on biomarker patterns of severe sepsis and septic shock. Critical

57. Vincent JL, Sakr Y, Sprung CL, et al. Sepsis Occurrence in Acutely Ill Patients Investigators: Sepsis in European intensive care units: results of the

SC

SOAP study. Critical Care Medicine 2006, 34: 344-353.

58. Boyd JH, Forbes J, Nakada TA,et al. Fluid resuscitation in septic shock: a

M AN U

positive fluid balance and elevated central venous pressure are associated with increased mortality. Critical Care Medicine 2011, 39: 259-265. 59. Murphy CV, Schramm GE, Doherty JA, et al. The importance of fluid management in acute lung injury secondary to septic shock. Chest 2009, 136:

TE D

102-109.

60. Vincent JL. The relationship between oxygen demand, oxygen uptake, and oxygen supply. Intensive Care Medicine 1990; 16:145-148. Marik PE, Baram M, Vahid B. Does central venous pressure predict fluid

EP

61.

AC C

responsiveness? A systematic review of the literature and the tale of seven mares. Chest 2008, 134:172-178.

62. Shoemaker WC, Appel PL, Kram HB, et al. Prospective trial of supranormal values of survivors as therapeutic goals in high-risk surgical patients. Chest 1988, 94: 1176-1186. 63. Hayes MA, Timmins AC, Yau EHS, Palazzo M, Hinds CJ, Watson D. Elevation of systemic oxygen delivery in the treatment of critically ill patients. New England Journal of Medicine 1994; 330:1717-1722.

28

ACCEPTED MANUSCRIPT 64. Alhasemi JA, Cecconi M, Hofer CK. Cardiac output monitoring: an integrative perspective. Critical Care 2011; 15:214. 65. Marik PE, Monnet X, Teboul JL. Hemodynamic parameters to guide fluid therapy. Annals of Intensive Care 2011; 1: 1-9.

RI PT

66. Michard F, Alaya S, Zarka V, Bahloul M, Richard C, Teboul JL. Global enddiastolic volume as an indicator of cardiac preload in patients with septic shock. Chest 2003; 124:1900-1908.

SC

67. Marik PE, Cavallazzi R, Vasu T, Hirani A. Dynamic changes in arterial waveform derived variables and fluid responsiveness in mechanically

M AN U

ventilated patients. A systematic review of the literature. Critical Care Medicine 2009, 37: 2642-2647. 68.

Cannesson M, Desebbe O, Rosamel P, et al. Pleth variability index to monitor the respiratory variations in the pulse oximeter plethysmographic waveform

TE D

amplitude and predict fluid responsiveness in the operating theatre. British Journal of Anaesthesia 2008, 101: 200-206. 69. Cavallaro F, Sandroni C, Marano C et al. Diagnostic accuracy of passive leg

EP

raising for prediction of fluid rsponsiveness in adults: systematic review and

AC C

meta-analysis of clinical studies. Intensive Care Medicine 2010; 36: 14751483.

70.

Saugel B, Ringmaier S, Holzapfel K et al. Physical examination, central venous pressure, and chest radiography for the prediction of transpulmonary thermodilution-derived hemodynamic parameters in critically ill patients: A prospective trial. Journal of Critical Care 2011; 26:402-410.

71. Goepfert MS, Richter HP, Eulenvurg CZ et al. Individually optimized hemodynamic therapy reduces complications and length of stay in the

29

ACCEPTED MANUSCRIPT intensive care unit: a prospective, randomized controlled trial. Anesthesiology 2013; 119:824-836. 72. Cordemans C, De laet I, Regenmortel NV, et al. Fluid management in critically illness patients: the role of extravascular lung water, abdominal

RI PT

hypertension, capillary leak, and fluid balance. Annals of Intensive Care 2012; 2: S1.

73. Sakka SG, Klein M, Reinhart K, Hellman AM. Prognostic value of

SC

extravascular lung water in critically ill patients. Chest 2002; 122: 2080-2086. 74. Jozwiak M, Silva S, Persichini R, et al. Extravascular lung water is an

M AN U

independent prognostic factor in patients with acute respiratory distress syndrome. Critical Care Medicine 2013; 41: 1-8.

75. Benes J, Chytra I, Altmann P, et al. Intraoperative fluid optimization using stroke volume variation in high risk surgical patients: results of prospective

TE D

randomized study. Critical Care 2010; 14: R118.

76. Wenkui Y, Ning L, JianFeng G, et al. Restricted peri-operative fluid administration adjusted by serum lactate level improved outcome after major

EP

elective surgery for gastrointestinal malignancy. Surgery 2010; 147: 542-552.

AC C

77. Valée F, Vallet B, Mathe O et al. Central Venous-to-arterial Carbon Dioxide Difference: an Additional Target for Goal-Directed Therapy in Septic Shock. Intensive Care Medicine 2006; 34;2218-2225.

78. Futier E, Robin E, Jabaudon M et al. Central Venous-to-arterial Carbon Dioxide Difference as complimentary tools for goal-directed therapy during high-risk surgery. Critical Care 2010; 14:R193.

79. Trzeciak S, McCoy JV, Dellinger RP et al. Early increases in microcirculatory perfusion during protocol-directed resuscitation are associated with reduced

30

ACCEPTED MANUSCRIPT multi-organ failure at 24 h in patients with sepsis. Intensive Care Medicine 2008; 34:2210-2217.

RI PT

Fig. 1. Frank–Starling curves are influenced by ventricular contractility. There is preload reserve when the ventricle is functioning on the steep part of the curve. This indicates preload responsiveness, where pulse pressure variation (PPV),

SC

stroke volume variation (SVV) and pulse variability index (PVI) are high, and end-expiratory occlusion (EEO) and passive leg raise (PLR) tests are positive.

M AN U

Volume loading induces a significant increase in stroke volume, and results in a small increase in extravascular lung water (EVLW). When the ventricle is functioning near the flat part of the curve, there is no preload reserve. This indicates preload unresponsiveness, where PPV, SVV and PVI are low, and EEO

TE D

and PLR tests are negative. Volume loading has little effect on stroke volume and

AC C

EP

leads to a large increase in EVLW. (Reproduced with Permission from [65])70

31

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Guiding principles of fluid and volume therapy.

Fluid therapy is a core concept in the management of perioperative and critically ill patients for maintenance of intravascular volume and organ perfu...
327KB Sizes 6 Downloads 9 Views