Cardiopulmonary Bypass and Management


THIS CHAPTER REVIEWS THE equipment and strategies for cardiopulmonary bypass (CPB) in infants and children, focusing on how they differ compared with CPB in adults. We review the effects of CPB on the key organ systems and discuss specific management issues that occur in daily practice.

Basic Aspects of Cardiopulmonary Bypass

The basic principles of CPB remain unchanged from when they were first introduced in the 1950s: the CPB machine assumes the functions of the heart and lungs during the time necessary to complete either an intracardiac or an extracardiac repair. A basic bypass circuit ( Fig. 19.1 ) consists of an oxygenator, heat exchanger, and venous reservoir; pump heads for perfusion, cardiotomy suction, and cardioplegia; and appropriate tubing, cannulas, and monitoring and alarm devices. Major differences exist between pediatric and adult CPB, stemming from anatomic, metabolic, and physiologic differences in these age groups ( Table 19.1 ).

FIGURE 19.1, Schematic diagram of a cardiopulmonary bypass circuit. This scheme depicts a membrane oxygenator with integral hard-shell venous reservoir and external cardiotomy reservoir. Many circuits have the cardiotomy reservoir, venous reservoir, and oxygenator integrated into one single unit. The systemic blood pump may be either a roller or centrifugal pump. Most pediatric venous cannulations are bicaval with two separate venous cannulas instead of the single venous cannula depicted here. Carbon dioxide can also be added to the inspired gas to facilitate pH-stat blood gas management. Arrows indicate direction of flow; P , pressure sensor; T , temperature sensors; X , placement of tubing clamps.

TABLE 19.1
Comparison of Pediatric Versus Adult Cardiopulmonary Bypass
Child Adult
Hemodilution 3–15 × adult Moderate
Perfusion pressure 30–40 mm Hg Moderate (>50–80 mm Hg)
Wide flow rates (0–200 mL/kg per minute) Narrow range (CI 2.0–2.4 L/m 2 per minute)
Blood gas management pH-stat (P co 2 20–80 mm Hg or greater) α-stat (P co 2 30–45 mm Hg)
Cannulation techniques Variable Predictable
Aortopulmonary collaterals Uncommon
Temperature ranges Variable DHCA occasionally
Glucose management Predictable
Inotropic response Negative Positive
Perfusion circuit Per kilogram weight Standard
Parameters Hematocrit often >55%–60%
P o 2 40–80 mm Hg ±
Sa o 2 75%–85%
Ultrafiltration (MUF/CUF) ± Ultrafiltration
CI, cardiac index; CUF, conventional ultrafiltration; DHCA, deep hypothermic circulatory arrest; MUF, modified ultrafiltration.

The Circuit and Cannulas

Unfortunately, the circuit size cannot be reduced proportionately to the patient's size; this disproportion commonly leads to hemodilution and dilutional coagulopathies in children. Surgical procedures require extremes of temperature, hemodilution, and changes in flow rates. Because of the smaller vascular structures and greater flow rates (150–200 mL/kg per minute) in infants and children compared with flow rates of 2.2 to 2.4 L/minute per meter squared in adults, selection of appropriately sized cannulas is critical to maintain these flows. Shear stress is significant in small cannulas and several-fold greater than needed for activation of blood cells and platelets, leading to a disproportionately exaggerated systemic inflammatory response syndrome (SIRS).

Bypass Circuit

Technical advances in the field of oxygenator construction and size, including the reduction of priming volumes to as small as 45 mL for neonatal oxygenators, have allowed marked reductions of circuit volumes over the past decade. Also, tubing sizes can be reduced to -inch diameters, which in combination with shorter length tubing, can allow the reduction of priming volumes to the range of 100 to 150 mL for neonates. A summary of the Texas Children's Hospital sizing chart is shown in Table 19.2 .

TABLE 19.2
Cardiopulmonary Bypass Circuit Prime Volume and Constituents at Texas Children's Hospital
Patient Weight Prime Volume Prime Constituents
<8 kg 350 mL Whole blood or PRBC + FFP + crystalloid prime a
8–15 kg 650 mL 100 mL albumin 25% ± PRBC + crystalloid prime a
15–25 kg 900 mL 100 mL albumin 25% + crystalloid prime a
15–25 kg 1200 mL Crystalloid prime a
FFP, fresh frozen plasma; PRBC, packaged red blood cells.

a Crystalloid prime: normal serum + PlasmaLyte + CaCl 2 + KCl.

Heparin-Coated Versus Noncoated Circuits

Young children are more susceptible to the adverse effects of CPB than adults, and the inflammatory response to CPB may have serious consequences for neonatal and pediatric patients. This is in part related to the surface area of the CPB circuit, which is large relative to the infant and child's blood volume. For example, a 3-kg neonate with a blood volume of 90 mL/kg has a total blood volume of approximately 270 mL, and with an average priming volume in many centers of 350 mL (120% of the neonate's estimated blood volume), the CPB circuit volume thus causes greater than 100% dilution. A 70-kg adult with 70-mL/kg blood volume has an approximately 5000-mL blood volume, and with a CPB circuit prime of 1500 mL, this results in less than 33% dilution. Contact of blood with the surface of the circuit also plays an important role for activation of coagulation and fibrinolysis. Heparin-coated biocompatible bypass systems reduce this activation in children weighing less than 10 kg undergoing CPB. They also reduce the activation of factor XII and the complement system. This results in less production of kallikrein and bradykinin, which in turn reduces the secretion of tissue plasminogen activator from endothelial cells. One study has documented more bleeding with a conventional, non–heparin-coated circuit compared with a heparin-coated circuit. Overall, children who underwent surgery while supported with heparin-coated circuits have significantly less inflammatory mediator release and fewer consequences thereof, such as prolonged postoperative ventilation and duration of stay in the intensive care unit (ICU).

Cardiopulmonary Bypass Pumps

The two pumps used most commonly for CPB are roller pumps and centrifugal pumps. Roller pumps have the advantages of simplicity, low cost, ease and reliability of flow calculation, and the ability to pump against increased resistance without reducing flow. Disadvantages include the need to assess occlusiveness, spallation or fragmentation of the inner tubing surface (potentially producing particulate arterial emboli), potential for pumping large volumes of air, and ability to create large positive and negative pressures. Compared with roller pumps, centrifugal pumps offer the advantages of less air pumping potential, less ability to create large positive and negative pressures, less blood trauma, and virtually no spallation. Disadvantages of centrifugal pumps include a greater cost, the lack of occlusiveness (creating the possibility of accidental patient exsanguination), and afterload-dependent flow that requires constant flow measurement. In the setting of short-term CPB for cardiac surgery, it remains uncertain whether the selection of a roller pump over a centrifugal pump, or of any specific centrifugal pump over another, has clinical importance. Pulsatile perfusion may prove to be beneficial in the future, but further outcome data and technical improvements are needed.

Cardiopulmonary Bypass Prime

The optimal priming fluid in cardiac surgery is a topic of enduring debate. Crystalloid solutions, colloids, and mixtures of both are used. Children appear to benefit from a colloid prime. If crystalloid is used for priming, it should not contain lactate or dextrose because CPB induces a metabolic acidosis that is iatrogenic, not splanchnic, in provenance. The addition of lactate to the prime increases the serum concentration of lactate postoperatively and should be avoided. Hyperchloremic metabolic acidosis is the second contributing component of a metabolic acidosis on CPB. This is often only detected by measuring the strong ion difference via the Stewart approach to the acid-base homeostasis. Both acidifying events are attenuated by the dilutional hypoalbuminemia induced by the pump prime. Because a hyperchloremic acidosis of a mild degree seems to be well tolerated and not associated with a poor outcome, no intervention seems necessary. Understanding the nature of CPB-associated acidosis, however, is likely to prevent unnecessary investigations or interventions.

The avoidance of dextrose is especially important during complex repairs using deep hypothermic cardiac arrest in which the risk of neurologic injury is substantive. The additives in banked blood, namely, glucose in citrate-phosphate-dextrose (CPD) storage solutions, also need to be considered as a source of glucose (together with the increased plasma concentrations of potassium in stored blood). We use a balanced electrolyte solution, such as PlasmaLyte, for the crystalloid component of our prime.

The proportionally large volume of the bypass circuit compared with the child's blood volume has a significant impact on the coagulation factors and cellular components. Platelet count decreases and coagulation factors, including fibrinogen, are diluted after bypass thereby contributing to a coagulopathy. The fibrinogen concentration at the end of bypass correlates with the 24-hour chest drainage in children who weigh less than 8 kg. This occurs more frequently in infants and neonates in whom the plasma concentrations of hemostatic proteins decrease postoperatively by 56% immediately upon initiation of bypass ; younger age represents the single most important risk factor for coagulopathy and bleeding complications.

One approach to offset this dilutional coagulopathy is the addition of whole blood to the circuit prime. Proponents cite two theoretical advantages: (1) improved hemostasis and (2) decreased SIRS with less edema formation and less organ dysfunction. However, one study challenged these perceived advantages when the researchers reported that the use of fresh whole blood actually increased perioperative fluid requirements, leading to a more prolonged duration of mechanical ventilation and ICU stay than in the single component group. The only advantage of whole blood prime was fewer donor exposures, a problem we obviate by matching packed red blood cells (PRBCs) and fresh frozen plasma (FFP) from the same donor. One retrospective analysis of donor exposures using fresh whole blood concluded that donor exposures were reduced in patients younger than 2 years of age compared with published reports using component therapy. Unfortunately, fresh whole blood is frequently unavailable. However, when a commitment to use fresh whole blood for pediatric cardiovascular surgery exists, it is possible to build a sustainable operating protocol to provide this resource. An alternative approach is to use FFP in the prime. Some investigators determined that the use of FFP led to greater fibrinogen concentrations at the end of surgery. On average, children in the FFP group needed 1.3 fewer donor exposures and tended to need fewer PRBCs. The reduced donor exposure was primarily the result of fewer transfusions of cryoprecipitate. FFP may be safely substituted by 5% albumin in the prime in children with less complex repairs and acyanotic lesions. Whenever possible, we prefer fresh blood that is less than 5 days old. Fresh PRBCs are presumably more balanced metabolically than stored PRBCs; the former contain less potassium, greater concentrations of glucose, reduced concentrations of lactate, and a greater pH. Also, postoperative morbidity increases with increasing age of red blood cells. Pulmonary complications, acute renal failure, and increased infection rates were among the main complications associated with increased red blood cell storage time. As far as potassium concentrations and acid-base balance are concerned, PRBC priming can be safely performed with stored PRBCs if the priming solution is circulated for 20 minutes before the initiation of CPB.

Depending on the size and age of the child and the complexity of the repair, a target hematocrit is chosen. Based on the child's blood volume and the prime volume, homologous blood is added using the following calculation:


Prime PRBC volume ( mL ) = [ Target hematocrit ] × [ Patient blood volume ( mL ) ] + Prime volume ( mL ) ] [ Patient PRBC volume ( mL ) ]

The average prime volume of the circuits in use at Texas Children's Hospital is shown in Table 19.2 . Other prime additives are heparin, antifibrinolytics, antiinflammatory agents (corticosteroids), antibiotics, vasodilators, and, sometimes, diuretics (mannitol, furosemide). At the end of the case and before separation from bypass, a blood gas sample should be analyzed to ensure the electrolytes (including calcium and magnesium ions), glucose, and hematocrit are within a desired range. Acid-base changes and sodium concentration are corrected with sodium bicarbonate and wash solutions, and residual lactate is washed out using hemofiltration.

Antifibrinolytic Agents

Aprotinin

Inhibitors of serine proteases regulate and prevent uncontrolled activation of thrombin, coagulation factors, complement products, kallikrein, trypsin, elastase, and cathepsin among others of these potent enzymes (see Chapter 20 ). Of the serine protease inhibitors, the broad-spectrum agent aprotinin is the most widely studied in both experimental and clinical settings. Aprotinin is derived from bovine lung. It inhibits plasmin, kallikrein, trypsin, and other proteases, resulting in both antiinflammatory and antifibrinolytic effects and maintenance of glycoprotein homeostasis.

The first use of aprotinin in pediatric cardiac surgery was reported in 1990 ; a high-dose regimen was administered to 28 children that included those undergoing a reoperation or surgery for transposition of the great arteries or endocarditis. No reduction in blood loss or drainage was observed; there were no adverse effects, and the time to chest closure from the end of cardiopulmonary bypass was reduced.

Despite the expense of aprotinin, follow-up studies reported more favorable results. Its use has reduced overall costs, from a reduced number of blood products used, operative time, duration of postoperative ventilation, and hospitalization. This was confirmed in a comparative analysis among antifibrinolytic medications. However, this benefit was observed only in complex repairs and the use of a high-dose regimen. The lesser effect of a low-dose regimen may be attributable to the dilutional effects in pediatric surgery compared with the adult population. Pediatric lung transplantation has been studied as a potential target group for the use of aprotinin. As in most high-risk groups, a significant benefit was found for children with repeat operations (defined as repeat sternotomies or repeat transplantations), either with a high- or a low-dose regimen. This is consistent with our experience. Also, in general, infants younger than 6 months of age and those with repeat sternotomies benefit from a high-dose regimen of aprotinin compared with reduced doses, despite greater drug costs. Economic studies have shown a cost-effective benefit of aprotinin in repeat cardiac procedures.

Aprotinin influences the inflammatory response to CPB in children. There has been a decrease in the duration of postoperative mechanical ventilation and an improved Pa o 2 /F io 2 (ratio of arterial oxygen concentration to the fraction of inspired oxygen, or P/F ratio), as an indicator of an attenuated reperfusion injury of the lung. The clinical relevance of its antiinflammatory action remains unclear but points toward significant antiinflammatory properties.

Although a standard dosing regimen has yet to be defined in children, pediatric studies have demonstrated decreases in operative time after CPB, in exposure to donor blood products, and in postoperative chest tube drainage. In vitro plasma concentrations of aprotinin have been related to antifibrinolytic and antiinflammatory activity at concentrations of 50 to 125 kallikrein inhibitor units (KIU)/mL and 200 KIU/mL, respectively. Anaphylactic and anaphylactoid reactions may occur with aprotinin, and a test dose should be given before administration of the loading dose or addition of aprotinin to the CPB circuit. In a retrospective review of 681 children, reactions occurred in 1% of first exposures, 1.3% of second exposures, and 2.9% of more frequent exposures.

We used aprotinin for complex neonatal repairs, such as arterial switch operations or Norwood procedures, as well as for most reoperative procedures and organ transplantations. The drug is currently unavailable in the United States and Europe because of safety concerns in adults, who presented with a different profile of complications after cardiac surgery than children. Aprotinin has been shown to be safe and effective in the neonate. Furthermore, serious questions have been raised regarding the statistical method used in the sentinel study that questioned the safety of aprotinin. Aprotinin continues to be used in Australia and New Zealand and has been reintroduced for adult coronary artery bypass graft surgery in Canada. Our dosing regimen is based on a 60,000 KIU/kg loading dose by the intravenous (IV) route and in the pump prime. The aprotinin infusion (7000 KIU/kg per hour) is started before skin incision. This infusion rate maintains the blood concentrations until the end of surgery at which time it is discontinued, just before leaving the operating room. Regimens based on body surface area are also used, along with a CPB prime dose that is based on the priming volume designed to achieve a plasma level above 200 KIU/mL. An example of one such calculation is a 0.85 to 1.7 × 10 6 KIU/m 2 loading dose both into the patient and the bypass prime, and an infusion of 2.0 to 4.0 × 10 5 KIU/m 2 per hour.

The Lysine Analogs: Aminocaproic Acid and Tranexamic Acid

Despite meticulous surgical technique, it is still frequently difficult to achieve adequate hemostasis after CPB, particularly in neonates. ε-Aminocaproic acid (EACA) and tranexamic acid (TXA) are analogs of the amino acid lysine that exert their antifibrinolytic effects by interfering with the binding of plasminogen to fibrin, thereby preventing the activation of the active plasmin (see Chapter 20 ). TXA may also improve hemostasis by preventing plasmin-induced platelet activation. Both EACA and TXA exercise some antiinflammatory properties, but not to the same extent as aprotinin. In one study, EACA reduced bleeding postoperatively in 25 of 71 children undergoing cardiac surgery on CPB but only benefited children with cyanotic heart disease. The empirical EACA loading dose was 75 mg/kg followed by an infusion of 15 mg/kg per hour, with an additional 75 mg/kg was added to the CPB prime. A larger loading dose of 150 mg/kg that was followed by an infusion of 30 mg/kg per hour of EACA has also been studied. In the latter case, intraoperative blood loss was reduced, although postoperative blood loss did not differ between the treatments. Blood coagulation measured with a thromboelastograph showed less fibrinolysis with EACA. The clearance of EACA is reduced in neonates compared with children and adults; dosing requirements in neonates were approximately half of those for children and adults. A regimen of 40 mg/kg as a loading dose, 30 mg/kg per hour infusion, and a pump prime concentration of 100 mg/L effectively maintained the plasma concentration in excess of 50 mg/L in 90% of neonates undergoing cardiac surgery using cardiopulmonary bypass. Dosing regimens for EACA at Texas Children's Hospital are displayed in Table 19.3 . A recent meta-analysis established the efficacy of EACA in pediatric cardiac surgery.

TABLE 19.3
ε-Aminocaproic Acid Dosing (Texas Children's Hospital)
Child dose from Ririe DG, James RL, O'Brien JJ, et al. The pharmacokinetics of epsilon-aminocaproic acid in children undergoing surgical repair of congenital heart defects. Anesth Analg. 2002;94(1):44–49.
Age (Weight) Loading Dose to Patient Infusion Dose Loading Dose to Bypass Circuit
<30 days (3.5 kg) 40 mg/kg 30 mg/kg per hour 0.1 mg/mL of CPB prime volume
1 months–12 years (3.5–40 kg) 75 mg/kg 15 mg/kg per hour 75 mg/kg
>12 years (>40 kg) 5 g 1 g/hour 5 g
Neonatal dose from Eaton MP, Alfieris GM, Sweeney DM, et al. Pharmacokinetics of epsilon-aminocaproic acid in neonates undergoing cardiac surgery with cardiopulmonary bypass. Anesthesiology 2015;122(5):1002–1009.

TXA compares favorably with EACA but confers a particular benefit in children with cyanotic heart disease alone. Those with acyanotic defects and those who required repeat sternotomies did not benefit from TXA, although that dosing regimen only included a single 50 mg/kg loading dose before incision. In children, the TXA plasma concentration between the peak of the loading dose and the end of CPB decreased 80% when it was not followed by a continuous infusion.

Although less efficient than aprotinin, EACA and TXA are equally effective in reducing perioperative blood loss in pediatric cardiac surgery. Given their safety profile, they may be even more appealing in the future. Further studies are needed to delineate their pharmacokinetic profiles and their efficacy. We use EACA based on simulation results from a study in children and adults. An initial loading dose of 75 mg/kg over 10 minutes followed by an infusion rate of 15 mg/kg per hour complemented a 75-mg/kg dose in the pump to maintain serum concentrations in excess of the therapeutic concentration (assumed to be 130 µg/mL) in more than 95% of children.

Special Coagulation and Hematologic Problems

Heparin-Induced Thrombocytopenia

The use of unfractionated heparin for anticoagulation for CPB in adults produces antiheparin antibodies in 25% to 50% of patients within 10 days postoperatively. In a small minority of these patients, high-titer immunoglobulin G (IgG) platelet-activating antibodies form and make immune complexes with heparin and platelet factor 4 (PF4). This results in activation of platelets (via their Fc receptors) and formation of procoagulant platelet microparticles, leading to thrombin generation and thrombosis. Thus the major problem in heparin-induced thrombocytopenia (HIT) is thrombocytopenia that occurs several days after heparin exposure accompanied by thrombosis, often in major vessels or structures. HIT appears to be less common, of milder course, and probably underrecognized in neonates and children. About 1% of children exposed to CPB have PF4 antibodies when tested before their second exposure to CPB, and actual HIT is much less common. When HIT is suspected, either PF4 enzyme-linked immunosorbent assay or a functional assay for HIT can be used to make the diagnosis; if positive, no further heparin should be given. If CPB is necessary, alternatives to heparin, such as the direct thrombin inhibitors argatroban, lepirudin, and bivalirudin, may be used. None of these agents is approved for use in children for anticoagulation for CPB, but case reports and small series have documented their successful use when HIT is diagnosed. The partial thromboplastin time (PTT), activated clotting time (ACT), and a specialized clotting time called the ecarin clotting time can be used to follow anticoagulation with these agents, but there is no reversal agent for them. Thus, treatment of post-CPB bleeding involves only administration of blood products and coagulation factors.

Antithrombin III Deficiency

Heparin produces anticoagulation by combining in a 1 : 1 ratio with antithrombin III (ATIII), which then binds to and inhibits thrombin, leading to anticoagulation. Of adult patients, 4% to 13% have a resistance to normal doses of heparin for CPB; most instances occur because of a partial deficiency of ATIII, rendering heparin less effective at producing anticoagulation. In children this is often unknown, and the first suspicion of ATIII deficiency may occur when the standard heparin dose of 300 to 400 units/kg fails to adequately anticoagulate before CPB; that is, the ACT remains less than 300 seconds. The usual response is to apply another dose of heparin from a different vial and remeasure the ACT, but if the ACT is still not adequately prolonged, a diagnosis of ATIII deficiency may be suspected. Infants younger than 6 months of age and children with congenital heart disease have decreased ATIII concentrations. Therefore heparin may not achieve adequate anticoagulation, and disorders in hemostasis and thrombosis and an exaggerated inflammatory response may occur. In this case, blood can be sent for ATIII levels, but to proceed with CPB, the ATIII must be increased. This can be accomplished in two ways: (1) by supplementing ATIII with 75 units/kg of recombinant ATIII and ensuring that the ACT is adequately prolonged before proceeding with CPB, or (2) by adding FFP (which has ample concentrations of ATIII) to the CPB prime or administering it to the child before bypass.

In the presence of reduced baseline concentrations of ATIII, the total dose of heparin, the amount of thrombin that was generated during bypass, and the fibrinogen that was consumed and fibrinolysis that was produced increased in infants undergoing cardiac surgery with CPB. This can exacerbate after bypass coagulopathy and transfusion requirements.

Recombinant Factor VIIa for Massive Hemorrhage

Recombinant factor VIIa (rFVIIa) was originally approved for use in patients with hemophilia who possess inhibitors to factors VIII or IX, and was shown to be effective in treating bleeding in these patients with doses of 90 µg/kg (see also Chapter 10 ). Endogenous factor VII circulates at small concentrations in the plasma. At a site of tissue or blood vessel injury, tissue factor (TF) is exposed, and the extrinsic coagulation pathway is activated by the binding of factor VII to TF, resulting in the activation of factor X to factor Xa, leading to the generation of thrombin from prothrombin, with further activation of platelets and the coagulation cascade. Large concentrations of rFVIIa activate the extrinsic pathway at the site of injury, theoretically without inducing systemic hypercoagulability. However, thrombotic complications are increased after its use. rFVIIa also activates platelets, adding to the potential benefit of this agent in significant hemorrhage. Thus this therapy seems appropriate for the treatment of surgical bleeding; a very complete review of the off-label uses of rFVIIa in pediatric cardiac surgery patients found no evidence to support the routine or prophylactic use of the therapy. However, rFVIIa may be beneficial as a rescue therapy for severe life-threatening refractory bleeding; the authors caution against the use of rFVIIa in children at risk for thromboembolic complications. A dose of 45 to 90 µg/kg, repeated every 2 hours, has been used. rFVIIa cannot produce hemostasis alone and should only be administered after the transfusion of sufficient amounts of platelets, plasma, and fibrinogen to form the substrate for hemostasis.

Fibrinogen Concentrate

Fibrinogen concentrate is an alternative to cryoprecipitate to replace fibrinogen after CPB. Compared with cryoprecipitate, fibrinogen concentrate, which is lyophilized and purified human plasma fibrinogen, has an improved safety profile because it has undergone viral inactivation and is devoid of microparticles that can cause vasoreactivity. In addition, fibrinogen concentrate contains a known amount of fibrinogen, and cryoprecipitate varies in the amount of fibrinogen per unit. One randomized pilot trial found no difference in the safety and efficacy of fibrinogen concentrate compared with cryoprecipitate when managing bleeding in children undergoing CPB. The dose of fibrinogen concentrate may be empiric (70 mg/kg), based on either the laboratory fibrinogen concentration or thromboelastometry.

Sickle Cell Disease

Sickle cell disease (SCD), one of the most common hemoglobinopathies among patients of African American or West Indian origin (with a prevalence of 0.2%–0.3% in that population), is the result of the substitution of valine for glutamic acid in position 6 of the β-hemoglobin chain. Normal adult hemoglobin is referred to as HbA, whereas hemoglobin containing the mutant β-hemoglobin chains is referred to as HbS. SCD is represented by a homozygous genotype (HbSS) with fractional concentrations of HbS in the range from 70% to 90%. Sickle cell trait, on the other hand, is a heterozygous manifestation (HbAS) with a prevalence of 8% to 10% in the same population. The definitive diagnosis of any sickle cell hemoglobinopathy is confirmed by hemoglobin electrophoresis (see Chapter 10 ).

Children with SCD are at particular risk for perioperative complications. Sickling can be triggered by hypoxia, dehydration, acidosis, hypothermia, stress, and infections. Hypoxia opens a Ca 2+ -activated K + channel (Gardos channel) that causes intracellular dehydration. Chain formation occurs that leads to increased blood viscosity with vasoocclusion. Opening of the Gardos channel, which is an important mechanism of sickle cell dehydration, depends on temperature, with greater potassium efflux at reduced temperatures. Shrinkage of sickle erythrocytes may also result from activation of a K + /Cl cotransport pathway under acidotic conditions. Activation of this pathway can be blocked by increasing the abnormally low level of intracellular magnesium in sickled erythrocytes. The use of magnesium and hydroxyurea in the perioperative period therefore seems to be reasonable.

CPB, particularly for more complex surgical procedures, may involve periods of low flow or even circulatory arrest, as well as hypothermia with consequent local vasoconstriction, hypoxemia, and acidosis. There is some evidence that CPB can be safely undertaken in SCD. Flow conditions are an important determinant of sickle erythrocyte adherence to endothelium. Under low-flow conditions, the adhesion of sickled cells to endothelium increases with contact time in the absence of endothelium activation or adhesive proteins, whereas under low-flow conditions in venules, sickle cell adhesion occurs only after endothelial activation. During CPB, both low-flow conditions and endothelial activation may occur. Multiple triggers of sickling are likely to occur during CPB, and close attention should be paid to the conduct of all aspects of bypass.

In the past, routine exchange transfusion has been recommended to prevent these complications. More recent experience provides evidence that not all children require an exchange transfusion. The growing evidence of the harmful effects of blood transfusion adds to the need to carefully reconsider routine exchange transfusion. For uncomplicated bypass surgery without periods of cardiac arrest, the omission of exchange transfusion has led to good outcomes.

Guidelines have been proposed for the perioperative management of children with sickle cell disorders. It is essential to avoid hypothermia using tepid or warm CPB in its stead; blood transfusion only for a decrease in hematocrit to less than 20%; maintenance of intravascular volume and body temperature while on CPB; the avoidance of vasopressors; the use of postoperative multimodal pain therapy; and early incentive spirometry to prevent pulmonary complications. In our practice, we use cerebral near-infrared spectroscopy (NIRS) to help determine an acceptable lower limit of hemoglobin for the individual child.

For children undergoing hypothermia, successful management with and without partial or complete exchange transfusion on bypass has been reported. Exchange transfusion can be performed preoperatively or on initiation of CPB. For exchange transfusion during CPB, the extracorporeal circuit is primed with blood and the usual components. When CPB is commenced, the child's blood volume is drained into storage bags and separated. The platelet-rich plasma is reinfused at the end of CPB, and the concentrated sickle cells are discarded. Platelet and plasma sequestration in conjunction with exchange transfusion reduces the need for postoperative transfusion and protects the platelets from the negative effects of CPB.

There seems to be no consensus as to a suitable target concentration of HbS. Reducing the absolute level of HbS may provide a greater benefit than targeting a particular ratio of HbA to HbS because the remaining sickle cells are still 100% at risk for sickling. In SCD, exchange transfusion has been shown to favorably affect cerebral tissue oxygenation. Exchange transfusion decreases both the proportion and absolute amount of HbS, but it does not remove every cell that may sickle. It may also improve hypoxic pulmonary vasoconstriction. In this context, these children may benefit from continuous hemofiltration to reduce inflammatory mediators and improve pulmonary recovery. Inhaled nitric oxide also has been recommended as an adjunct to prevent sickle cell crisis. It may improve the binding of oxygen, thereby reducing the formation of sickle cells; reduce pulmonary hypertension; and improve pulmonary function without adverse effects on normal hemoglobin.

A Perspective on Blood Preservation: Cardiopulmonary Bypass in Jehovah's Witness Patients

Jehovah's Witnesses differ from other religious groups in their conscious objection to the therapeutic infusion of blood and blood components. They uniformly refuse the transfusion of red blood cells, and some individuals also refuse platelets and plasma, as well as predonated autologous blood. Individual choices that can be made are the acceptance of fractions of blood, such as albumin and globulins, dialysis, cell savage, and acute isovolemic hemodilution (see Chapters 5 and 12 ).

Acute isovolumic reduction of hemoglobin to a concentration of 5 g/dL has been well tolerated in healthy children under anesthesia in one study and does not appear to reduce tissue oxygenation significantly. Reduction of oxygen delivery to 7 to 8 mL/kg per minute under resting conditions does not increase the oxygen debt. This degree of anemia is compensated for, in part, by an increased extraction, an increase in cardiac index, and a subsequent decrease in systemic vascular resistance. In a retrospective study of the morbidity associated with reduced concentrations of hemoglobin in Jehovah's Witness patients, the hemoglobin concentration of those who died was less than 5 g/dL. A safe limit of hemodilution in children has not been established. Hemodilution in acyanotic children up to 50% appears to be well tolerated and safe, although in cyanotic children, hemodilution probably should not exceed 40%. If this level of hemodilution is exceeded, hemodynamic instability and inadequate oxygen delivery can occur. Evidence suggests that hematocrit concentrations of 21.5% in infants on CPB significantly increase adverse psychomotor developmental outcomes compared with concentrations of 27.8%.

The most important and simplest strategy to avoid transfusion in the setting of cardiac surgery is to limit blood loss. Unnecessary and reduced amounts of blood removed for testing and sampling reduce the blood loss. Pharmacologic agents, such as aprotinin, TXA and EACA, reduce the risk of perioperative blood loss. The administration of erythropoietin in the cardiac surgery setting has been shown to reduce the risk of exposure to allergenic blood. Preoperative recombinant erythropoietin is an acceptable strategy to Jehovah's Witnesses to augment the red cell concentration. This strategy requires that oral iron (2–6 mg/kg of elemental iron in 2–3 divided doses) and vitamin C are started about 6 weeks before surgery followed by twice weekly erythropoietin (50–100 IU/kg) intramuscularly about 3 weeks before surgery. Hemoglobin concentrations should be tracked to ensure that the concentration does not exceed 15 to 20 g/dL as venous thromboembolism may occur. Some Jehovah's Witnesses refuse albumin, a constituent in the preparation of erythropoietin that is supplied in glass ampoules. In the latter case, a lyophilized preparation of erythropoietin, which is albumin-free, may be used. The cost of erythropoietin can be substantial, and one cost analysis suggested that its use in cardiac surgery is not cost-effective.

Intraoperative recovery of blood with a cell salvage device is also acceptable to many Jehovah's Witnesses. This involves the removal by suction of blood from the operative field followed by washing, filtering, and return of red blood cells to the patient. A randomized controlled trial of intraoperative cell salvage in cardiothoracic surgery demonstrated a reduction in RBC transfusion and an increase in postoperative hemoglobin.

Acute normovolemic hemodilution involves the preoperative removal of a volume of blood from the patient with the simultaneous administration of crystalloid or colloid to maintain circulating volume. The collected blood is then reinfused during the operation. Some Jehovah's Witnesses find this process acceptable, especially if the blood remains in continuity with the patient throughout. Acute normovolemic hemodilution has other advantages, including lower costs, because the blood does not need compatibility testing; reduced the possibility of administrative error; and achieved patient time-saving (see also Chapters 10 and 12 ). The development of artificial red cell substitutes could potentially obviate the need for compatibility testing, as well as vastly reduce infection risks, with none of the immunomodulatory side effects of allogeneic blood. Some of these products would also be acceptable to Jehovah's Witness families. Substitutes include perfluorocarbons, hemoglobin solutions, intramolecular cross-linked hemoglobin, and liposome-encapsulated hemoglobin. None of these has reached clinical practice. Lastly, autologous retrograde priming has been used in Jehovah's Witness patients and can further reduce the hemodilutional effects of the prime. For this purpose, priming of the arterial line of the CPB circuit is accomplished with the patients' own blood.

Modern bypass circuits reduce the priming volumes to less than 200 to 300 mL. Main components that are amenable to volume reduction on a regular circuit are the size and length of the lines, small oxygenators and arterial filters, and priming the hemofilter for modified ultrafiltration with blood from the venous line after CPB. Line volumes, for example, may vary from 1.73 mL per 10 cm of a -inch tubing to 0.75 mL per 10 cm of a -inch tubing. The limiting factor, however, is the necessary flow. For a -inch arterial line, a maximum flow of 1.8 L/minute was established as the point at which the Reynolds number reaches a critical value indicating turbulent flow, which may damage RBCs. Modified ultrafiltration at the end of CPB through a fluid warmer line to prevent heat loss or continuous ultrafiltration has been used. The venous line and the reservoir are emptied before discontinuation of bypass, the field is suctioned, and all blood is retransfused through the arterial line. Decannulation is achieved and protamine is given as usual. Crystalloid cardioplegia solution should be evacuated from the field by an external sucker to prevent dilution of the pump volume.

Postoperative care involves minimal blood sampling, and only on special indications. Noninvasive monitoring allows uncomplicated weaning from the ventilator. The first report of successful outcomes in Jehovah's Witness children with congenital cardiac defects was in 1985 ; 110 children older than 6 months of age successfully underwent operation, with a perioperative mortality rate of 5.3%. Only one death was attributed to blood loss. A weight less than 5 kg is considered by some as a contraindication for open-heart surgery and palliative procedures were advocated in the past. For some lesions, however, no palliation is possible. The development of miniaturized circuits, preoperative optimization, use of antithrombolytic drugs, vacuum-assisted drainage to allow smaller tubing and cannula sizes, as well as the use of modified ultrafiltration, enabled the safe expansion of surgery into the neonatal population. Individualized heparin level–based anticoagulation management further results in a reduction of coagulation problems, blood loss, and transfusion requirements. The addition of desmopressin, 0.3 µg/kg, is thought by some to improve platelet activity and stimulate the release of von Willebrand factor after protamine infusion, although this is not evidence-based.

In one study, when center-specific blood conservation strategies were used, bloodless cardiac surgery was most successful in children greater than 18 kg in weight, followed by those 6 to 18 kg in weight. All 73 patients less than 6 kg in weight had transfusions during their hospitalization.

All of the aforementioned considerations are important in approaching the Jehovah's Witness patient; however, at Texas Children's Hospital, Jehovah's Witness children are not treated differently with regard to blood transfusion practice than any other child. Cerebral NIRS is used to help determine the safe hemoglobin level for the individual child at all phases of surgery. Consent for blood transfusion in this situation is a complicated issue, because the legal status of children differs from that of an adult. Each institution must develop a legal informed consent process for blood transfusion for Jehovah's Witness children, in consultation with local legal authorities, social work and ethnic groups, and representatives of the Jehovah's Witness faith (see Chapter 5 ). Currently, we have a release of liability form for the parents to sign stating that he or she requests that no blood products be used, but acknowledges they may be needed to treat his or her child. The parent further agrees to release and hold harmless the physicians and hospital for any liability associated with blood transfusion. This form was developed in conjunction with the local Jehovah's Witness church representatives, and in our practice this has been accepted by more than 95% of parents and has obviated the need for more extreme measures, such as temporary child protective services custody during the perioperative period, which was our former practice.

Myocardial Protection

Myocardial protection during cardiac surgery has evolved over the years, and the concept of chemical cardioplegia was introduced in 1955. Before the popular use of chemical cardioplegia, topical cardiac hypothermia was used. In the late 1970s and early 1980s, the concept of cold hyperkalemic blood cardioplegia was introduced. Potassium concentrations in cardioplegic solutions ranging from 12 to 30 mEq/L are typically used to achieve cardiac standstill within 1 to 2 minutes under hypothermic conditions, with greater concentrations or induction times required for normothermic conditions. Myocardial edema after bypass and global ischemia can be reduced by a number of strategies that involve modifying the conditions of delivery and composition of cardioplegia solutions as they affect the movement of intracellular and interstitial fluid. In contrast to studies in adults, most studies conducted in neonates have shown little difference between blood and crystalloid cardioplegia. Hypothermia also decreases myocardial oxygen consumption. The benefits of this approach appear to be optimal at myocardial temperatures between 24°C and 28°C. However, there is growing evidence that warm, intermittent blood cardioplegia may be advantageous to either cold crystalloid or cold blood cardioplegia. The benefits of blood cardioplegia are more pronounced in younger, cyanotic children who require longer aortic cross-clamping. For acyanotic children, the cardioplegic technique is probably not as critical. Avoidance or reduction of myocardial edema occurs by limiting the pressure of cardioplegia infusions and by providing moderately hyperosmolar cardioplegia solutions that contain blood. Buffering the acidosis that results from ischemia is achieved by including tromethamine, histidine-imidazole, or both in the cardioplegia solution. Close management of myocardial calcium balance to avoid extremes of intracellular hypercalcemia or hypocalcemia, especially during reperfusion, is very important. The addition of magnesium may solve this dilemma by preventing damage from greater cardioplegic calcium concentrations by its action as a calcium antagonist. This prevents mitochondrial calcium overload as a consequence of reperfusion injury. Magnesium also prevents the influx of sodium into the postischemic myocardium, which is exchanged for calcium during reperfusion.

Every cardiac program has its own philosophy regarding cardioplegia and myocardial protection. At Texas Children's Hospital, plain crystalloid cardioplegia is used. The prime blood gas and electrolytes should mimic physiologically the child's arterial blood gas as closely as possible. If whole blood or packed cells are added to the prime, the target hemodilution range should be 28% to 30%; the prime should be recirculated continuously and warmed between 35.0°C and 36.5°C before initiation of bypass. In neonates and infants, albumin is added to the cardioplegic solution to maintain an appropriate colloid osmotic pressure. This may decrease edema formation of the arrested heart. In children undergoing circulatory arrest, long cross-clamp times, and large pump suction return cases, 20 mg/kg methylprednisolone is used, up to a maximum of 500 mg, to reduce the production of inflammatory mediators that result in myocardial dysfunction. Table 19.4 summarizes the Texas Children's Hospital protocols for cardioplegia and myocardial protection.

TABLE 19.4
Cardioplegia Solution
CARDIOPLEGIA BASE SOLUTION (385 ML)
Concentration Contents
Sodium chloride BP 3.54 g/L Sodium 23 mmol
Anhydrous glucose BP 6.65 g/L Potassium 15 mmol
Potassium chloride 2.92 g/L Calcium 0.35 mmol
Mannitol 6.54 g/L Chloride 39 mmol
Calcium chloride 135 mg/L Glucose 2.52 g
Mannitol 2.48 g
Approximate pH 4.5
275 mOsm/L
CARDIOPLEGIA BUFFER SOLUTION
Concentration Contents
Sodium carbonate 9.37 g/L Sodium carbonate 0.28 g
Sodium bicarbonate 27.0 g/L Sodium bicarbonate 0.81 g
USES OF CARDIOPLEGIA SOLUTION DURING CARDIOPULMONARY BYPASS
Children Weighing <10 kg
385 mL Cardioplegia base solution
26 mL Cardioplegia buffer solution
100 mL 25% Albumin
Note: This is usually delivered at a pressure of 30 mm Hg for newborns and 30–40 mm Hg for older infants.
Children weighing >10 kg
385 mL Cardioplegia base solution
100 mL 0.9% Sodium chloride
10 mL 25% Mannitol
5 mL 8.4% Sodium bicarbonate
Note: This is usually delivered at a pressure of 30–60 mm Hg. A good guide is to note the end-diastolic pressure of each child before bypass. This will be a guide to the normal filling pressure of the coronary arteries. When aortic incompetence is present, the CPS flow may need to be increased.
ADMINISTRATION OF CARDIOPLEGIA SOLUTION
For All Patients:
Temperature 8°C–12°C
Initial dose 110 mL/m 2 per minute for 4 minutes
Subsequent doses 110 mL/m 2 per minute for 2 minutes
Note: Following the initial dose, cardioplegia is delivered every 20 minutes during the cross-clamp period unless otherwise indicated by the surgeon. The perfusionist will remind the surgeon of the need for cardioplegia and keep track of the time. Because of the nature of the surgical procedure, it may be necessary to deliver cardioplegia directly into the coronary ostia via a handheld delivery system. In this case, the surgeon will direct the perfusionist. Close attention should be paid to the delivery line pressures.
EXAMPLES OF PRIMES
Neonate: Whole Blood, if Available, Otherwise Reconstituted
Whole blood 225 mL
PlasmaLyte A 50 mL
0.45% NaCl 125 mL
Heparin 2500 units
NaHCO 3 5 mEq
CaCl 2 250 mg
Pediatric: Packed Red Blood Cells
PRBCs 250 mL
Plasmalyte A 300 mL
0.45% NaCl 75 mL
25% Albumin 100 mL
Heparin 3500 units
NaHCO 3 20 mEq
CaCl 2 300 mg
Adult: Crystalloid Prime
PlasmaLyte A 700 mL
0.45% NaCl 600 mL
25% Albumin 100–200 mL (volume varies depending on the size of the patient)
5% Dextrose 40 mL
Heparin 5000 units
NaHCO 3 40 mEq
CaCl 2 300 mg
KCl 2.4 mEq
BP, the material conforms to the specifications and procedures outlined in the British Pharmacopoeia; CPS, cardioplegia solution; PRBCs, packed red blood cells.

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here