The Toyota approach to anaesthesia- small continuous improvements: using placebo, IV cannulation, echo, blocks and compression devices

Toyota is famous for improving their cars through a process of continuous, small, incremental improvements, a technique known as Kaizen, or the Toyota way. In this way many small improvements, each inconsequential on their own, when added together produce significant results.

I think this is a great model to use when looking at anaesthesia. Anaesthesia and surgery are complicated processes, and most of the “low hanging fruit” in terms of safety improvements have already been made. It is unlikely that any single factor will make a major difference to outcomes. However that doesn’t mean we should stop trying to improve, and using a wide range of small improvements in different areas will collectively improve the patient’s experience.

An example of this is IV cannulation, something we all do every day and which we often forget can be quite painful. In addition, this is often patient’s only way of judging the ability of their anaesthetist. I recall a poster presentation where only 2 things determined a patients satisfaction with their anaesthetist- did the IV hurt, and did they visit the patient more than just in recovery. The patient has no baseline to judge postoperative pain or nausea and can awaken after all sorts of intraoperative near-catastrophes none the wiser, however if you want to make a patient happy, get the drip in first time and make sure it doesn’t hurt!

Two articles on this topic have got my attention and changed my practice. Both involve randomised trials where patients were either warned they were going to feel a “sharp sting” or used more neutral and comforting words, eg  “I am going to apply the tourniquet on the arm. As I do this many people find the arm becomes heavy, numb and tingly. This allows the drip to be placed more comfortably”. The patients not only reported lower pain scores but were also less likely to withdraw their hand in the “kind words” group compared to the “nocebo” group. This is contrary to the common practice of warning someone, with the rational that it then won’t be as bad as they expect. In fact all this achieves is heightened anxiety and more pain (read here and here).

This is an example of avoiding the “nocebo” (opposite of placebo) effect of harsh words like sharp, sting, needle and pain. There is increasing evidence that placebo plays a major part in many interventions. Recently i went to an intriguing talk about placebo where the concept of the “open/hidden” trial was discussed. This is the opposite to a placebo controlled trial. Instead of everyone getting told they were getting morphine and half getting a sugar pill, all patients get given morphine but only half are told about it. The rest had it quietly slipped into a bag of fluid without being told. There were significantly greater reductions in pain in the group that were told they were getting the “powerful painkiller”, compared to the group that had it slipped into their fluids. The presenter gave a range of slides for different analgesics showing that for virtually all of them the pain score reductions were double in the open  “powerful painkiller” group compared to the hidden ones.

Finally, three further topics for the “continuous improvement” theme, all of which i will talk about more in the future.

The first is transthoracic echo for use by anaesthetists in preoperative assessment. This is something that was big when I was doing my fellowship at the Royal Melbourne Hospital and is spreading around the world rapidly. In this month’s Anaesthesia the RMH team have provided the first (weak) evidence that preoperative echo may improve outcome, instead of simply changing management (which has been shown in previous studies). The study is observational, of poor quality, subject to the Hawthorne effect and shows an implausibly large mortality difference, but for all that makes pleasing reading for transthoracic echo exponents such as myself (reference here).

The second is the use of dexamethasone to prolong peripheral nerve blocks. This is something we have been doing recently in our hospital in Mackay in Australia, and the results can only be described as “spectacular, bordering on scary” – 24-30 hours duration from a single shot interscalene block, including complete motor block at 24 hours. This is consistent with studies showing dexamethasone effectively doubles the duration of most nerve blocks. Just remember that the phrenic nerve is also paralysed for 24 hours!

When I first read about this I had some concerns on potential neurotoxicity, but these were alleviated by 2 things.  The first of these were the words of a chronic pain physician colleague who stated that they add dexamethasone to every block they do, have been doing so for years and have had no problems. The second is a study showing that in an animal model dexamethasone was significantly less neurotoxic that ropivicaine, and the ropi/dex combination was less toxic than other common combinations such as ropi / buprenorphine and ropi/ clonidine (reference here and here and here)

The final interesting note is on SCDs- the sequential calf and thigh compressors now ubiquitous on the legs of patients having surgery in our hospitals. Two recent articles showed that they reduce intraoperative hypotension- a bonus that at first seems unexpected until you think about it, then seems quite logical.

Reduced hypotension for caesarian:  here

This study used a variation of the normal compressors with higher pressures and longer compression times: here

Agitation and Pain in the Recovery Room (tutorial)

Problem:

A 43 year old male returns from the operating room following cholecystectomy. The operation had been originally planned using the laparoscopic approach. However it became necessary to convert to an open procedure. Intraoperatively the patient received fentanyl 300mic/g, propofol, vecuronium, oxygen and desflurane and cefazolin. At the end of surgery, neuromuscular blockade (sustained tetanus was demonstrated) was reversed, the patient opened his eyes and was extubated.

On arrival to the recovery room the patient is combative, thrashing around, incoherent, not obeying commands, attempting to remove has urinary catheter. His pulse rate is 120 beats per minute, blood pressure 170/100, temperature 37.0 degrees Celcius and his pulse oximeter is reading an SpO2 of 99%.

1. Identify the problem

What is the mechanism of injury and what are the treatment options?

This patient is agitated: the most common cause of postoperative agitation is pain. Pain is a neurohormonal and emotional response to a noxious stimulus, in this case surgical injury. Pain is the “fifth vital sign.”

Pain is known to worsen perioperative outcomes: it results in – increased protein catabolism – thereby reducing physiologic reserve, retention of salt and water, impaired wound healing, prolonged recumbent times (resulting in increased risk of deep venous thrombosis), and significant suffering and dissatisfaction on the part of the patient. Elevated adrenergic activity results in increased oxygen demand and may precipitate myocardial ischemia. In patients, such as in this case, that undergo upper abdominal surgery, the splinting effect of pain results in impaired coughing and lung derecruitment and increased risk of pulmonary complications including nosocomial pneumonia.

One of the major roles of perioperative clinicians is to minimize patient suffering. Patients universally report dissatisfaction with perioperative pain management.1 Modern approaches to preventing suffering in perioperative patients include a multimodal approach to pain, postoperative nausea and vomiting, anxiety, agitation and delirium.2-5

2. Understand the problem

What is pain (understanding the mechanisms)?

Table 1.  Inflammatory Mediators that amplify the pain response
Substance
Source
Norepinephrine Nerve endings & circulating
Epinephrine Circulating
Substance P Nerve endings
Glutamate Nerve endings
Bradykinin Plasma kininogen
Histamine Platelets, mast cells
Hydrogen Ions (acidity) Ischemia / Cell Damage
Protaglandins Arachidonic acid / damaged cells
Interleukins Mast Cells
Tumor necrosis factor alpha Mast Cells

Surgical incision is associated with tissue injury and release of inflammatory mediators, development of local edema and activation of nocioceptors. These are nerve endings of myelinated (A-delta) and unmyelinated (C) afferent nerve fibres that respond to noxious thermal, mechanical, or chemical stimulation. A-delta fibres are mechanothermal while the C fibres are polymodal.

When nociceptors are activated, a series of neurohormonal reflexes are activated, and a painful sensation is elicited.6 In the awake patient, this is apparent by an adverse emotional response, a sensation of “unpleasantness”. In a sedated patient this may result in hyperadrenergic activity, agitation or aggressiveness.

Figure 1: Pain Pathways

1. The nocioceptive response is activated at the level of the surgical incision; 2. release of inflammatory cytokines and vasodilator metabolites; 3 transmission of nocioceptor impulses along afferent A-delta and C fibers; 4. integration and amplification in the spinal cord – c”windup”; 5 transfer of impulses from doral horns to thalamus and post-central gyrus; 6. activation of hypothalmo-pituitary adrenal axis; 7 release of cortisol, epinephrine and norepinephrine; 8 central and peripheral sensitization .

Normally, there is a relatively high threshold for activating nocioceptors. However, tissue injury alters the activity of these neurones, due to the  local production of inflammatory mediators (an “inflammatory soup” table 1). These include substance P, glutamate, bradykinin, histamine and arachadonic acid metabolites, such as prostaglandins. Their impact is twofold, to directly activate nocioceptors, and to reduce the firing threshold of these receptors.7 This is traditionally known as peripheral sensitization: lower stimuli than usual result in pain sensation. This is amplified in part by the systemic production of catecholamines secondary to activation of the hypothalmo-pituitary-adrenal axis. Moreover, epinephrine and norepinephrine induce a state of anxiety and diaphoresis that worsens the emotional response (figure 1). In addition, tissue trauma activates inflammation, and inflammation causes pain. Inflammation causes pain through the up-regulation of stimulated nociceptors and the recruitment of nonstimulated or dormant receptors.8 Proinflammatory mediatiors such as interferons, tumor necrosis factor alpha, interleukin-1 and interleukin 6 decrease the threshold for impulse generation and increase the intensity of the nocioceptive response.7 A patient emerging from anesthesia with elevated levels of stress hormones (cortisol and catecholamines) that is experiencing significant pain, will frequently become agitated and inappropriate, as in this clinical scenario.

Nocioceptor activation results in all-or-nothing depolarization of the afferent nerve. Painful impulses are transmitted to the dorsal horn spinal cord and subsequently to the thalamus and the post central gyrus.  In the spinal cord central sensitization may occur.9 A sufficiently strong stimulus may change the interpretation of painful impulses and subsequent stimuli are amplified. An area of hyperalgesia, composed of undamaged tissue, may appear adjacent to the injured site.10 This is due to “windup”, which results from repetitive C-fiber stimulation, mediated by glutamate via n methyl d aspartate (NMDA) receptors.11

Second order neurones synapse at the level of the spinal cord and transmit pain signals to the brain. Two predominant types of second order neurones have been identified: wide dynamic range (WDR) neurones and nociceptive specific (NS) neurones. Nociceptive signals ascend in the spinothalamic and spinoreticular tracts. These fibres project to multiple sites in the brain stem and midbrain, including the brain stem autonomic regulatory sites, hypothalmus and thalmus.

The body does have a pain regulating system that attenuates the response. This is modulated by a variety of neurotransmitters and inhibitory interneurones, that utilize endogenous encephalins and endophins and gamma aminobutyric acid (GABA). The binding of these endogenous opioids to central and peripheral receptors results in reduced presynaptic release of neurotransmitters, in particular substance P, and curtailed nocioceptor response.7

MANAGING PAIN – TREATMENT OPTIONS

Opioids

Opioids remain the mainstay of treatment for postoperative analgesia. Opioids exert their effects by binding to an array of receptors (“opioid receptors” μ, κ and δ) that exist in the central and peripheral nervous system and gastrointestinal tract. This results in analgesia and an array of characteristic side effects (table 2). In addition, opioids may have anti-inflammatory and immunomodulatory effects.7 A variety of naturally occurring, synthetic and semi-synthetic opioid agents are available for therapeutic use. These include full μ receptor agonists, partial agonists and agonists/antagonists (table 3). The choice of agent is dependent on the practice patterns of the clinician The majority of us use limited selection of full opioid receptor agonists, including morphine, hydromorphone, fentanyl, oxycodone, meperidine and methadone. Tramadol and codeine are weak receptor agonists.

Table 2: Side effects of Opioids
Pruritus
Nausea and Vomiting
Sedation
Dysphoria
Respiratory Depression
Urinary retention
Constipation
Hypotension
Bradycardia
Urticaria
Confusion

Physicochemical Properties of opioids

There are two physiochemical properties of opioids that determine their pharmacologic action (table 4): degree of ionization and lipid solubility.  Opioids are weak bases. When dissolved in solution, they are dissociated into protonated and free-base fractions, with the relative proportions depending on the pH and pKa. The more unionized the agent, the more rapid its onset of action (table 3): hence alfentanil, which is 80% unionized (it has a pKa of 6.1) and remifentanil which is 70% unionized (pKa 7.1,) have more rapid onset of action than fentanyl (<10% unionized, pKa 8.4).

Lipid solubility is determined by the chemical structure of the agent. The more lipid soluble the agent the more easily the agent passes thru the blood brain barrier to the site of action. This also impacts onset of action: hence fentanyl has a more rapid onset of action than sufentanil which is more rapid thanmorphine or hydromorphone. Lipid solubility also impacts the volume of distribution of the agent: the higher the lipid co-efficient, the greater the amount of the drug sequestered in fat stores in the body. This is important when opioids are used as continuous infusions, resulting in a complex pharmacologic process known as “context sensitive half time.”

After intravenous injection, arterial plasma concentrations of opioids rise to a peak within one circulation time. Thereafter, they exhibit a rapid redistribution phase and a slower elimination phase typical of drugs whose pharmacokinetics are described by multi-compartmental models. Drugs that are more lipid soluble, such as fentanyl and sufentanil, redistribute extensively to fat, including non receptor fatty tissue in the brain. Alfentanil and remifentanil, agents that are not lipid soluble, have low volume of distribution and are rapidly cleared from plasma. Morphine, hydromorphone and meperidine have relatively low lipid solubility and are extensively metabolized by the liver. They have relatively long duration of action. Fentanyl and similar agents are also extensively cleared by the lungs.

Table 3: Opioid Agents available of Postoperative Analgesia

Agent

Dose im/iv

Oral Dose

Morphine 10mg equal to:

Onset (min) iv

Peak Effect (min)

Duration

Alfentanil

500mg

NA

300mg

<1

1-2

10-20 min

Buprennorphine

0.3mg

0.5mg

2.5mg

15

60

5 hr

Butorphanol

1-4 mg

1 mg

2 mg

5-10

45

3-4 hr

Codeine

103 mg

200 mg

130 mg

30

60

3-4 hr

Dezocine

5-10mg

NA

10 mg

15-30

60-90

3-4 hr

Fentanyl

100 mg

NA

125 mg

0.5

5

1 hr

Hydromorphone

2 mg

4mg

1.3 mg

15-30

2-3

30-60 min

Levorphanol

2 mg

2-3 mg

2.3 mg

25

45

6-8 hr

Meperidine

50-100 mg

100mg

75 mg

1- 2

30-60

2-4 hr

Methadone

2.5-10 mg

2.5-10 mg

10 mg

30-60

30-60

4-6 hr

Morphine

10 mg

30-60 mg

10 mg

2-3

20

4 hr

Nalbuphine

10 mg

NA

12 mg

2-3

30

4-6 hr

Oxycodone

NA

5 mg

10 mg

15-30 (PO)

60

4 hr

Oxymorphine

1 mg

NA

1.1

1-2

30-60

4-6 hr

Pentazocine

30 mg

50 mg

60 mg

15-30

100

3-4 hr

Propoxyphene

NA

200 mg

200 mg

15-30 (PO)

120

4 hr

Sufentanil

20 mg

NA

12.5 mg

<1

2

30-45 min

 

Morphine

Morphine, a component of opium, has been used for analgesia and anxiolysis for millennia. If was first purified by Serturner, a German pharmacist, in 1803. He called this alkaloid “Morphia” after Morpheus, the Greek God of Dreams. Morphine is a phenanthrene opioid receptor agonist that exerts its major effects on the CNS and gastrointestinal tract. It is the prototype opioid analgesic agent, against which all other agents are compared. The majority of healthcare professionals are familiar with this drug in terms of its clinical effects, dosing and complications. This imparts a significant degree of safety; as a result I recommend morphine as the first line analgesic agent in PACU.

Following injection the onset of analgesic effect of morphine is 5 minutes with a peak effect at 20 minutes. Morphine is predominantly unionized (pKa 8.0) and has low lipid solubility: penetration into the brain is consequently relatively slow (table 4). Patients experience mild sedation prior to analgesia. This makes morphine an ideal analgesic agent for patients that are agitated and in pain, such as in this scenario, and patients requiring mild sedation for mechanical ventilation in the PACU or ICU. Conversely, the neurologic assessment of patients with brain injuries of following neurosurgery may be clouded by this effect.

Single boluses of morphine may be ineffective to establish adequate analgesia. Aggressive “loading” with the drug, to break the cycle of pain, may be required. This requires careful titration to analgesic and sedative response (figure 2).

Table 4 :  Physicochemical and pharmacokinetic data of commonly used opioid agonists
  Morphine Meperidine Fentanyl Sufentanil Alfentanil Remifentanil
pKa 8.0 8.5 8.4 8.0 6.5 7.1
% Un-ionized at pH 7.4 23 <10 <10 20 90 67
Partition Co-efficient 1 32 955 1727 129 16
% Bound to plasma protein 20–40 39 84 93 92 80?
Elimination half time (hours) 1.7-3.3 3.0-5.0 3.0-6.6 2.2-4.6 1.4-1.5 0.17-0.33

Morphine is known to have a direct histamine releasing effect. While the clinical implications of this are generally overstated, transient vasodilatation and hypotension may result. This is an unlikely problem in hyperadrenergic postoperative patients complaining of pain, but may be an issue in patients under general anesthesia or who are receiving concomitant propofol infusions. Morphine may cause euphoria. It alters the perception of pain: the patient knows that he/she is in pain, but is not bothered by it.

Figure 2: Effect of administration recurrent boluses of morphine on pain, and sedation in a typical postoperative patient in PACU. VAS = visual analog score; RASS = Richmond agitation sedation scale.

Morphine has an important side effect profile (table 2): it is a direct respiratory depressant, and acts by reducing the respiratory center’s responsiveness to carbon dioxide. Morphine induces nausea and vomiting by an effect on the chemoreceptor trigger zone. It induces miosis (pupillary constriction), causes constipation, may cause urinary retention and causes cutaneous vasodilatation. Morphine is a potent anti-tussive agent; again this may be beneficial in mechanically ventilated patients. Morphine is principally metabolized by conjugation in the liver, to morphine 3-glucuronide (M3G), which is inactive and  morphine 6-glucuronide (M6G), which is highly potent. There is also some extrahepatic metabolism in the kidney. Hence care should be taken when morphine is administered to patients with significant renal impairment, as delayed respiratory depression may follow.

Hydromorphone

Hydromorphone is structurally very similar to morphine. It differs from morphine by the presence of a 6-keto group and the hydrogenation of the double bond at the 7-8 position of the molecule.12 It principally acts at µ receptors, and thus shares a similar side effect profile. Hydromorphone is slightly more lipid soluble than morphine, and has a slightly quicker onset of action; its peak effect is at 20 minutes. Hydromorphone is less sedating than morphine and does not have active metabolites (although it is metabolized by the liver and metabolites accumulate in renal failure). Hydromorphone is widely used for patient controlled analgesia and for intravenous analgesia in the ICU. The major limitation of using hydromorphone is confusion regarding the appropriate bolus dose. Hydromorphone is roughly 7.5 times more potent than morphine; as a result one is more likely to encounter accidental (due to prescription error) overdose with this agent.

Fentanyl

Fentanyl, alfentanil, sufentanil and remifentanil are semi synthetic opioids that have rapid onset and relatively short duration of action. Only fentanyl is routinely used for postoperative analgesia. It may be administered intravenously as bolus or infusion, transdermally through a patch or novel transcutaneous delivery systems, transorally (fentanyl “lollipop”), intrathecally or epidurally.13

Fentanyl has relatively rapid onset of action (1-2 minutes peak effect 5 mins) and short duration of action (20 minutes). However its therapeutic window is relatively narrow. When fentanyl is administered in low to moderate dose (1-5 mic/kg) intraoperatively, there may be little or no residual drug effect by the time the patient arrives in PACU. The patient may experience severe pain.

Fentanyl is highly lipophilic and redistributes to fat stores in the body; this may result in significant accumulation if given in high dosage. Its context sensitive half time is relatively long if administered by infusion. The complex pharmacology of fentanyl limits its effectiveness for perioperative analgesia. For prolonged effect, high dosage (5-15mic/kg) need to be administered, risking significant respiratory depression. Tachyphylaxis develops rapidly resulting in reduced effectiveness with escalating dose. This may result in significant problems such as ileus and urinary retention.

Although fentanyl can be used for analgesia in PACU, its effectiveness is limited to short episodes of analgesia, for example if coverage is required during a procedure – such as placement of a chest tube or epidural. It is not an effective agent for significant visceral pain unless given as an infusion or thru a PCA (patient controlled analgesia device). Transcutaneous fentanyl patches have slow onset of action and have no role in acute pain management. Newer products that utilize iontophoresis (a non-invasive method of propelling high concentrations of a charged substance transdermally using a small electrical charge), may make patient controlled fentanyl administration popular for ambulatory surgery.

Meperidine (Pethedine)

Meperidine is a semi-synthetic opioid structurally similar to fentanyl. Meperidine is one-tenth as potent as morphine. Meperidine is an effective analgesic and, in equianalgesic dosage, produces as much sedation, euphoria, respiratory depression and nausea and vomiting as morphine. Meperidine is significantly different pharmacologically to morphine, and has effects on a medley of receptors (see chapter on shivering). Of interest, meperidine has atropine like effects. The majority of opioids cause bradycardia, presumably by a direct or indirect action on the hypothalmo-pituitary-adrenal axis. Meperidine induces tachycardia. It also causes papillary dilatation. Meperidine has no anti-tussive effects. It has smooth muscle relaxing effects, and was used traditionally as analgesic for heptobiliary and ureteric surgery. However there is no evidence that this agent is superior to morphine in these situations. Meperidine causes less constipation and urinary retention than morphine. It has been used for generations for intramuscular analgesia in labor. The major current clinical use of meperidine is for treatment of postoperative anesthesia related shivering.

The major limitation of meperidine is its active metabolites: normeperidine (norpethidine) and meperidinic acid. Normeperidine accumulates, particularly in renal failure and may cause CNS stimulation (seizures or myoclonus)

Tramadol

Tramadol is an atypical opioid which is a centrally acting analgesic, used for treating mild to moderate pain. It is a synthetic agent, as a 4-phenyl-piperidine analogue of codeine.It can be administered orally or intravenously.

Tramadol is approximately 10% as potent as morphine, when given intravenously. It has effects on opioid, GABAergic, noradrenergic, NMDA (antagonism) and serotonergic receptors. Analgesia with tramadol is not fully reversed with naloxone although it has weak affinity for the μ-opioid receptor (approximately 1/6th that of morphine). The major issue with this agent is the serotonin modulating properties that may lead to interaction with selective serotonin reuptake inhibitors and result in serotonin syndrome (see chapter on malignant hyperpyrexia).

Tramadol causes significantly less respiratory depression and bowel dysfunction than conventional opioid analgesics.14 It does cause nausea and vomiting and may reduce seizure threshold.

Patient Controlled Analgesia

There is tremendous inter-patient variability in postoperative analgesic requirements. Coupled with greater demands on, and reduced availability of, nurses on postoperative wards, patient controlled analgesia has emerged as the gold standard delivery system for postoperative pain relief.15 Patients prefer PCA to nurse administered analgesia.16 Dolin and colleagues17 collected pooled postoperative pain scores from 165 publications and concluded that the mean incidence of moderate to severe pain was 67.2% and that of severe pain 29.1% for intramuscular opioids. The corresponding values were 35.8% and 10.4% for PCA, and 20.9% and 7.8% for epidural analgesia, respectively. The superiority of epidural analgesia has been confirmed by other investigators.18 Nonetheless, PCA, compared with conventional opioid treatment, improves analgesia and decreases the risk of pulmonary complications.19 In a large meta-analysis of fifty-five studies with 2023 patients receiving PCA and 1838 patients assigned to a control (parenteral ‘as-needed’ analgesia), PCA provided better pain control and greater patient satisfaction.20 Patients using PCA consumed higher amounts of opioids than the controls and had a higher incidence of pruritus (itching) but had a similar incidence of other adverse effects. There was no difference in the length of hospital stay.

Surprisingly, these results were not seen in many studies. This probably relates to the tremendous variability in settings applied to PCA devices: bolus doses, lockout, background infusions, opioid agents used etc.21 PCA strategy should be titrated to patient requirements. The best available evidence suggests that the optimal bolus dose of morphine is 1mg.22 Initial IV-PCA bolus doses of other drugs that are commonly used for opioid-naive patients are hydromorphone, 0.2 mg; fentanyl, 20 to 40 μg.21 The lockout interval is used to limit the frequency of demands made by the patient within a certain time. Lockout periods between 5 and 10 minutes are commonly prescribed. If analgesia is inadequate with a certain lockout period, it is more effective to increase the bolus dose rather than reducing the lockout.23 The use of a background infusion with IV-PCA, in addition to bolus doses on demand, is targeted at improving patient comfort and sleep: the expectation is that the patient will not awaken in pain. However, studies report no benefit to pain relief or sleep and no decrease in the number of demands made but a marked increase in the risk of respiratory depression.21 Background infusions should be limited to use in chronic opioid users.

Multimodal Analgesia

The concept of multimodal analgesia is based on the observation that pain is a multifactorial phenomenon – amplified and modulated at different sites both peripherally and centrally, and is therefore not amenable to control by opioid monotherapy alone.2;4;5

Figure 3: Multimodal Analgesia: a balanced approach to analgesia – different agents are combined to reduce pain transmission at different sites, targeting local nerves and neurotransmitters, the CNS and the neuroendocrine system.

The multimodal approach to perioperative pain management involves attenuating nociceptive activity at many different levels (Figure 3), including:

  • Reducing nocioception output at the surgical site, by superficial and deep wound infiltration.
  • Treating and preventing peripheral inflammation using nonsteroidal anti-inflammatory drugs (NSAIDs).
  • Blocking afferent nerve activity by regional blockade using local anesthetics with or without other agents (opioids, tramadol, and ketamine). This blockade may be involve peripheral nerves (inguinal field block) neural plexuses (brachial or lumbar plexus block) or  spinal level (subarachnoid or epidural block)
  • Modulating central pain processes at the level of the brain or spinal cord (e.g., opioids, tramadol, NMDA antagonists, alpha-agonists). Some agents, such as opioids, have been shown to work at a number of levels (peripheral, spinal, cerebral).
  • Reducing adrenergic activity by direct or indirect actions using opioids or alpha-2 adrenoceptor agonists.

Wound infiltration with local anesthetic

Patients recurrently complain of pain at the site of superficial injury, i.e. at the skin incision site. Local anesthetic infiltration of the surgical site may reduce this.  A number of approaches have been shown reduce postoperative analgesia requirements including: infiltration into the subfacial parietal peritoneum, subcutaneous infiltration and field block, intraarticular injection, drains lavage etc.    Wound infiltration is a safe, simple, effective and under-utilized postoperative analgesic technique.

Non Steroidal Anti-inflammatory Agents (NSAIDS).

Non steroidal anti-inflammatory agents (NSAIDS) have been used for some time in ambulatory surgery to reduce the dose of opioid required for pain relief, with the potential for less nausea, vomiting and sedation. NSAIDS act peripherally to inhibit the cyclo-oxygensase enzymes responsible for production of pro-inflammatory mediators at the site of injury. The major therapeutic limitation of using these agents is the delay between application and onset time, because of peripheral action, as compared with opioids.  In order to attain full efficacy, it is essential to give NSAIDS at induction or early during the procedure.  NSAIDS, whether COX-1 or COX-2 inhibitors, are equally efficacious in terms of pain relief but their clinical applications are limited by concerns relating to gastric bleeding, renal impairment and platelet dysfunction.2 In general NSAIDS should be avoided in patients that have undergone mucosal surgery (such as resection of the nasal mucosa), intracranial surgery, some spinal surgeries and operations in which a cross-clamp has been placed on the aorta (abdominal aortic aneurysm repair for example). In addition, patients with known renal dysfunction are at risk of acute renal failure, and NSAIDS should be withheld.

NSAIDS may be administered by a variety of routes including oral, intravenous, intramuscular, and rectal.  Agents routinely utilized in the operating room and PACU include ketoralac, diclofenac and tenoxicam. There is a considerable body of evidence supporting the use of NSAIDS as adjunct analgesic agents.2;8;24  For example, ketoralac 30mg has equipotent analgesic effect to morphine 10mg. Where possible patients should be administered NSAIDS in the operating room or PACU.

Regional anesthetic techniques

A multitude of different regional anesthetic techniques has been used for surgery. These are frequently combined with general anesthesia to ensure absence of pain in the postoperative period. For example, paravertebral block has emerged as an effective option for breast surgery in addition to general anesthesia.25 Combinations of ileoinguinal and ileohypogastric nerve blockade and caudal block, have been shown to significantly reduce postoperative pain in children and adults following inguinal hernia repair.26 The commonly used regional nerve blocks are featured in Table 5.

Table 5   Regional Blocks that may reduce Postoperative Pain
Block Type Indication
Upper-limb blocks:
Bier’s block Surgery to hand or wrist (e.g. Colles’ Fracture)
Digital Nerve Block Surgery to finger
Wrist Block: median, ulnar and radial nerves Surgery to hand
Elbow Block: median, ulnar and radial nerves Surgery to hand or wrist
Brachial Plexus Block:
Axillary Approach Surgery to hand, wrist or lower arm
Supra-calvicular Approach Surgery to hand, wrist, upper and lower arm
Interscalene Approach Surgery to upper limb and shoulder
Neuraxial Blocks:
Spinal Surgery to lower extremities
Epidural Surgery to lower extremities, abdomen, thorax
Caudal Surgery to perinuem (e.g. hemmoroidectomy)
Paravertebral Block Thoracic and abdominal surgery (e.g. breast surgery, herniorrhaphy).
Lower-limb Blocks:
Sciatic Nerve Block Surgery to lower limb
Obturator Nerve Block Surgery to lower limb
3-in-1 (Lumbar Plexus) Block Surgery to lower limb
Knee Block: common peroneal, tibeal & saphenous Surgery to lower leg
Ankle Block Surgery to foot
Truncal Blocks
Intercostal Blocks Thoracic surgery
Inguinal Field Blockade Surgery in lower abdomen (e.g. hernia repair)
Penile Block Surgery to penis (e.g. circumcision)

Alpha-2 receptor agonists.

The alpha2 -agonists clonidine and dexmedetomidine have been reported to provide effective analgesia following a variety surgical procedures and when given by oral, intrathecal and intravenous routes of administration.

In general, alpha 2 -agonists are best used as adjuncts with other analgesics to minimize the side effects of sedation and hypotension. Clonidine, when given orally as a premedication (5 mg/kg), reduces morphine PCA requirements by 37% and significantly reduces the incidence of nausea and vomiting.27 When added to local anaesthetics, clonidine has been shown to augment the effectiveness and duration of action of peripheral nerve blocks.28

Paracetamol/Acetaminophen

Paracetamol is an agent commonly used in multimodal techniques, due to its wide availability and low side effect profile in therapeutic dosage. Oral and rectal acetaminophen, as an adjunct to opioids, reduces pain scores by 20% – 30%.29 It has analgesic and antipyretic, but not anti-inflammatory, activity. Although the mechanism of action of acetaminophen is poorly understood, it is believed to act by the inhibition of the COX-3 isoenzyme and subsequent reduced prostanoid release in the central nervous system. In addition, there is some suggestion that it acts on the opioidergic system and NMDA receptors. Acetaminophen is a weak analgesic agent and has little or no anti-inflammatory activity. Thus it has no role as monotherapy-analgesia following major surgery. Nevertheless, there is abundant evidence that acetaminophen significantly enhances analgesia when combined with opioids and NSAIDS. It has little or no impact on the gastrointestinal tract or kidney. However, in high dosage acetaminophen may cause irreversible liver damage. This agent is strongly recommended for balanced analgesia in perioperative patients.

3. Differential diagnosis / Work the problem

What is the differential diagnosis?

An agitated patient, emerging from anesthesia, is in pain until otherwise proven (figure 4). It is imperative to assess the patient’s respiratory status to ensure that he is oxygenating and ventilating as hypoxemia and hypercarbia may manifest as agitation. The patient’s agitation should also be assessed in terms of their total physical status: in general agitation plus tachycardia and hypertension suggests hyperadrenergic activity (stress), and agitation associated with bradycardia suggests increased vagal activity. This can result, for example, from distress associated with a full bladder.

Postoperative patients that are agitated and tachycardic may have partial neuromuscular blockade (chapter 4: hypertension and tachycardia). Other potential diagnoses include drug withdrawal (beta blockers, clonidine, alcohol, cocaine and amphetamines), drugs (atropine, neostigmine, naloxone or flumazenil) and pathologic processes (neurologic injury, electrolyte abnormalities and endocrinopathy). Neurologic injuries include ermergence delirium, stroke, intracranial bleed, raised intracranial pressure and transcranial herniation. Electrolyte abnormalities that may cause agitation include hypernatremia, hyponatremia, hypokalemia, hypophosphatemia, hypercalcemia and hypomagnesemia). Endocrinopathies that may cause agitation include thyrotoxicosis, diabetic ketoacidosis, hypoglycemia, pheochromocytoma and carcinoid syndrome.

Figure 4: Managing the Agitated Patient

Common things are common – once life threatening causes of agitation have been outruled, one must address pain.

The clinical assessment of pain, analgesia, anxiety and sedation require quantification, hence the use of scoring systems. The behavioral pain score was discussed in chapter 4; it allows the bedside clinician determine whether or not the sedated patient is in pain. In this scenario the patient is agitated, and the level of agitation should be assessed using an alternative system, such as the Richmond Agitation Sedation Scale (RASS table 6).30 This tool allows the clinician to assess whether the patient is agitated or sedated using a + (plus) score for agitation and a minus (-) score for sedation. It then allows for titration of sedative drugs. The scale scores the patient from -5 (comatose) to +4 (combative, as in this case).

Visual Analog Scale for Pain

 No Pain           Mild Pain             Moderate Pain              Severe  Pain                  Worst

_____________________________________________________________________

0         1             2            3          4        5          6            7          8          9            10

Figure 5: Visual Analog Scale

Once the patient is co-operative, pain should be assessed using a visual analog scale (VAS figure 6).31;32 This is a 0-10 scoring system in which 0 is no pain and 10 is the worst pain the patient has ever experienced. The goal is to obtain a pain score of 3 or less or pain that is considered acceptable by the patient.33 Occasionally the patient may report a higher score than would be suspected by physiologic data, and the bedside nurse is required to make an objective decision about the need for further analgesia.

Table 6    Richmond Agitation Sedation Scale

Clinical Status

RASS

Combative (violent dangerous to staff)

4

Very agitated (pulling on or removing catheters)

3

Agitated (fighting ventilator)

2

Anxious

1

Spontaneously alert calm and not agitated

0

Able to maintain eye contact >= 10 seconds

-1

Able to maintain eye contact < 10 seconds

-2

Eye opening but no eye contact

-3

Eye opening or movement with physical or painful stimuli

-4

Unresponsive to physical or painful stimuli (deeply comatose)

-5

4. Solve or resolve the problem  

Step 1: Ensure that the airway is patent and that the patient is breathing spontaneously. Apply supplemental oxygen. Ensure that the patient has iv access and that intravenous fluid is running. Check the patient’s pulse and blood pressure. Position the patient in the semi-recumbent position.

Step 2: Score the patient’s agitation/pain using RASS and VAS

Step 3: Commence the opioid titration protocol (figure 6).34;35 The choice of agent, morphine or hydromorphone is determined by the clinician – if the patient is to receive a morphine PCA they should receive bolus morphine, a hydromorphone PCA – bolus hydromorphone etc. The dose should be adjusted for the patient’s weight.

Step 4: The goal of the opioid titration protocol is to aggressively treat pain by assessing the patient’s pain score,33 and to prevent oversedation by using the RASS score.

Step 5: If the patient remains agitated despite significant opioid administration, consideration should be given to anxiety and delirium. Delirium is defined as an acute disturbance of consciousness (reduced clarity of awareness of the environment) and cognition with reduced ability to focus, sustain or shift attention. The patient may be disorientated in time, place or person. If the patient is orientated, anxiety may be a problem, and this can be managed with judicious administration of midazolam 1-2mg iv. Anxiolytics should not be administered before analgesics in the agitated postoperative period unless the significant consideration has been given to pain as the etiology of the problem.

Step 6: If the patient is complaining of severe intractable pain, out of proportion to the injury and unresponsive to analgesia and anxiolysis, consideration should be given to an surgical problem. For example, in a patient that has had orthopedic surgery to the leg, severe pain may signal a compartment syndrome: the patient has ischemic pain. If a surgical cause had been discounted, consideration should be given to a regional anesthetic approach (epidural, brachial plexus catheter etc) or to the addition of ketamine to the PCA.

Step 7: If the patient becomes oversedated (RASS -3 or below) as a result of narcosis (associated with bradypnea), opioid administration should be discontinued until the RASS score is -2 or above. In extreme cases, where the patient is comatose and hypoventilating, naloxone should be administered in aliquots of 40mic/g, until the RASS is -2 or above. It is essential that the clinician consider alternative causes of coma, such as stroke, brain hemorrhage or intracranial hypertension).

Figure 6: Management of Patient in Pain or Agitated

Conclusions

  1. Pain is now considered the “fifth vital sign”.
  2. Postoperative patients that are agitated should be considered to be in pain until otherwise proven.
  3. Pain is a multisystem problem that manifests as an emotional response to a noxious stimulus. Pain starts at the nocioceptor and is amplified by local inflammatory mediators and spinal cord windup leading to central and peripheral sensitization. In addition, pain activates the hypothalmo-pituitary-adrenal axis leading to anxiety and diaphoresis.
  4. Pain should be managed by a multimodal approach that addressed the problem at different levels in the pain pathways.
  5. Opioid agents remain the mainstays of management of pain. Of these morphine and hydromorphone are the most popular and effective agents for managing visceral pain in PACU.
  6. Opioids should be titrated using an opioid titration protocol, that scores both pain and sedation.
  7. Anxiety, delirium and surgical problems may worsen pain, and should be addressed by the clinician.
  8. Pain that is unresponsive to aggressive and analgesic therapy should prompt the clinician to consider a surgical cause.

This PBLD was written by Patrick Neligan Version 1.3 September 2007

 

References

1.    Myles PS, Williams DL, Hendrata M, Anderson H, Weeks AM: Patient satisfaction after anaesthesia and surgery: results of a prospective survey of 10,811 patients. British Journal of Anaesthesia 2000; 84: 6-10

2.    Joshi GP: Multimodal analgesia techniques and postoperative rehabilitation. Anesthesiol.Clin.North America. 2005; 23: 185-202

3.    Kehlet H, Dahl JB: The value of “multimodal” or “balanced analgesia” in postoperative pain treatment. Anesth.Analg. 1993; 77: 1048-56

4.    Kehlet H: Multimodal approach to control postoperative pathophysiology and rehabilitation. Br.J.Anaesth. 1997; 78: 606-17

5.    White PF, Kehlet H, Neal JM, Schricker T, Carr DB, Carli F, the Fast-Track Surgery Study Group: The Role of the Anesthesiologist in Fast-Track Surgery: From Multimodal Analgesia to Perioperative Medical Care. Anesthesia Analgesia 2007; 104: 1380-96

6.    Brennan TJ, Zahn PK, Pogatzki-Zahn EM: Mechanisms of incisional pain. Anesthesiol.Clin.North America. 2005; 23: 1-20

7.    Cohen MJ, Schecter WP: Perioperative pain control: a strategy for management. Surg Clin.North Am. 2005; 85: 1243-57, xi

8.    Siddall PJ, Cousins MJ: Pain mechanisms and management: an update. Clin.Exp.Pharmacol.Physiol 1995; 22: 679-88

9.    Woolf CJ: Central sensitization: uncovering the relation between pain and plasticity. Anesthesiology 2007; 106: 864-7

10.    Wolpaw JR, Tennissen AM: Activity-dependent spinal cord plasticity in health and disease. Annu.Rev.Neurosci. 2001; 24: 807-43

11.    Herrero JF, Laird JM, Lopez-Garcia JA: Wind-up of spinal cord neurones and pain sensation: much ado about something? Prog.Neurobiol. 2000; 61: 169-203

12.    Murray A, Hagen NA: Hydromorphone. Journal of Pain and Symptom Management 2005; 29: 57-66

13.    Stanley TH: Fentanyl. Journal of Pain and Symptom Management 2005; 29: 67-71

14.    Desmeules JA: The tramadol option. Eur.J Pain 2000; 4 Suppl A: 15-21

15.    Lehmann KA: Recent Developments in Patient-Controlled Analgesia. Journal of Pain and Symptom Management 2005; 29: 72-89

16.    Ballantyne JC, Carr DB, Chalmers TC, Dear KB, Angelillo IF, Mosteller F: Postoperative patient-controlled analgesia: meta-analyses of initial randomized control trials. J Clin.Anesth 1993; 5: 182-93

17.    Dolin SJ, Cashman JN, Bland JM: Effectiveness of acute postoperative pain management: I. Evidence from published data. British Journal of Anaesthesia 2002; 89: 409-23

18.    Werawatganon T, Charuluxanun S: Patient controlled intravenous opioid analgesia versus continuous epidural analgesia for pain after intra-abdominal surgery. Cochrane.Database.Syst.Rev. 2005; CD004088

19.    Walder B, Schafer M, Henzi I, Tramer MR: Efficacy and safety of patient-controlled opioid analgesia for acute postoperative pain. A quantitative systematic review. Acta Anaesthesiol.Scand. 2001; 45: 795-804

20.    Hudcova J, McNicol E, Quah C, Lau J, Carr DB: Patient controlled opioid analgesia versus conventional opioid analgesia for postoperative pain. Cochrane.Database.Syst.Rev. 2006; CD003348

21.    Macintyre PE: Intravenous patient-controlled analgesia: one size does not fit all. Anesthesiol.Clin.North America. 2005; 23: 109-23

22.    Owen H, Plummer JL, Armstrong I, Mather LE, Cousins MJ: Variables of patient-controlled analgesia. 1. Bolus size. Anaesthesia 1989; 44: 7-10

23.    Macintyre PE: Safety and efficacy of patient-controlled analgesia. British Journal of Anaesthesia 2001; 87: 36-46

24.    White PF: The Changing Role of Non-Opioid Analgesic Techniques in the Management of Postoperative Pain. Anesthesia Analgesia 2005; 101: S5-22

25.    Karmakar MK: Thoracic paravertebral block. Anesthesiology 2001; 95: 771-80

26.    Nehra D, Gemmell L, Pye JK: Pain relief after inguinal hernia repair: a randomized double-blind study. Br.J.Surg. 1995; 82: 1245-7

27.    Park J, Forrest J, Kolesar R, Bhola D, Beattie S, Chu C: Oral clonidine reduces postoperative PCA morphine requirements. Can.J.Anaesth. 1996; 43: 900-6

28.    Eisenach JC, De Kock M, Klimscha W: alpha(2)-adrenergic agonists for regional anesthesia. A clinical review of clonidine (1984-1995). Anesthesiology. 1996; 85: 655-74

29.    Schug SA, Sidebotham DA, McGuinnety M, Thomas J, Fox L: Acetaminophen as an adjunct to morphine by patient-controlled analgesia in the management of acute postoperative pain. Anesthesia Analgesia 1998; 87: 368-72

30.    Ely EW, Truman B, Shintani A, Thomason JW, Wheeler AP, Gordon S, Francis J, Speroff T, Gautam S, Margolin R, Sessler CN, Dittus RS, Bernard GR: Monitoring sedation status over time in ICU patients: reliability and validity of the Richmond Agitation-Sedation Scale (RASS). JAMA 2003; 289: 2983-91

31.    Bodian CA, Freedman G, Hossain S, Eisenkraft JB, Beilin Y: The Visual Analog Scale for Pain: Clinical Significance in Postoperative Patients. Anesthesiology 2001; 95: 1356-61

32.    Aubrun F, Hrazdilova O, Langeron O, Coriat P, Riou B: A high initial VAS score and sedation after iv morphine titration are associated with the need for rescue analgesia. Can.J Anaesth. 2004; 51: 969-74

33.    Aubrun F, Langeron O, Quesnel C, Coriat P, Riou B: Relationships between measurement of pain using visual analog score and morphine requirements during postoperative intravenous morphine titration. Anesthesiology 2003; 98: 1415-21

34.    Aubrun F, Monsel S, Langeron O, Coriat P, Riou B: Postoperative titration of intravenous morphine. Eur.J Anaesthesiol. 2001; 18: 159-65

35.    Aubrun F, Monsel S, Langeron O, Coriat P, Riou B: Postoperative titration of intravenous morphine in the elderly patient. Anesthesiology 2002; 96: 17-23

 This Article Copyright Patrick Neligan MA MB FCAI DIBICM 2007-2012. Neither text nor illustrations are to be used without permission. 


Perioperative Single Dose Ketorolac to Prevent Postoperative Pain

Anesth Analg. 2012 Feb;114(2):424-33

I have been a sceptic of meta-analyses, for many years. The purpose of these studies is to take a large volume of literature; both published and unpublished, and test hypotheses. The idea is to replicate a large randomized controlled trial by combining many studies together.  Unfortunately, rather than coming up with hundreds of studies involving thousands of patients, invariably the selection criteria excludes a large number of studies, and relatively few are ultimately analyzed. Often no conclusions are reached or the results are refuted by a subsequent large  randomized controlled trial.

This month’s Anesthesia and Analgesia features a  meta-analysis of single dose ketorolac for prevention of postoperative pain. Ketoralac is a NSAID analgesic very similar to diclofenac – but importantly one that can be given by intravenous bolus (I routinely use it in my clinical practice). The analysis included 13 studies, approximately 800 patients: ketorolac versus placebo, for prevention of postoperative pain.

I was surprised to find that the most efficacious dose of ketorolac was 60mg – this dose provided effective opioid sparing analgesia and reduced the incidence of PONV (post operative nausea  and vomiting). Interestingly, intramuscular administration of ketorolac appears to be more effective than intravenous. The authors claimed that the standard dose of ketorolac, 30mg, was essentially ineffective.

Will this study change my clinical practice?

When reading studies such as this you really have to go beyond the abstract. With meta-analyses it is essential to see what question the authors were asking:

“We included only randomized clinical trials of a single perioperative (preoperative or intraoperative) systemic ketorolac with an inactive (placebo or “no treatment”) control group. Included studies had to report at least pain scores or opioid consumption on postoperative pain outcomes.”

Also “Trials evaluating >1 dose of perioperative ketorolac were also excluded to maximize clinical homogeneity. Studies containing a concurrent use of a different analgesic regimen were excluded if a direct comparison of ketorolac and placebo could not be established.”

In other words – if investigators had planned to give more than one dose of ketorolac or were using a multimodal strategy, their studies were excluded. In other words – my clinical practice was excluded.

The accompanying editorial, from Henrik Kehlet’s group makes several criticisms of the study (Click here) and is reassuring to those of us who use ketorolac in multimodal regimes. They systematically review the literature that includes ketorolac, clearly demonstrating efficacy of the 30mg dose. They also, and very helpfully, demonstrate that ketorolac has minimal association with postoperative complications.

What I learned from this paper:  single low dose NSAID is inadequate for postoperative pain relief.

ASA Periopeative Pain Guidelines

The ASA has issued guidelines for acute pain management in the perioperative setting. Click on this link to read guidelines. Obviously there is a North American flavour to these guidelines, but they are generalisable.

Summary:

1. Institutional Policies and Procedures for Providing Perioperative Pain Management

  • Anaesthetists offering perioperative analgesia services should provide ongoing education and training to ensure that hospital personnel are knowledgeable and skilled with regard to the effective and safe use of the available treatment options within the institution.
  • Educational content should range from basic bedside pain assessment to sophisticated pain management techniques (e.g., epidural analgesia, PCA, and various regional anesthesia techniques) and non-pharmacologic techniques (e.g., relaxation, imagery, hypnotic methods – these are US guidelines!).
  • For optimal pain management, ongoing education and training are essential for new personnel, to maintain skills, and whenever therapeutic approaches are modified.
  • Anaesthetists and other healthcare providers should use standardized, validated instruments to facilitate the regular evaluation and documentation of pain intensity, the effects of pain therapy, and side effects caused by the therapy.
  • Anaesthetists responsible for perioperative analgesia should be available at all times to consult with ward nurses, surgeons, or other involved physicians.
  • They should assist in evaluating patients who are experiencing problems with any aspect of perioperative pain relief.
  • Anaesthetists providing perioperative analgesia services should do so within the framework of an Acute Pain Service.
  • They should participate in developing standardized institutional policies and procedures.

2. Preoperative Evaluation of the Patient

  • A directed pain history, a directed physical examination, and a pain control plan should be included in the anaesthetic preoperative evaluation.

3. Preoperative Preparation of the Patient

  • Patient preparation for perioperative pain management should include appropriate adjustments or continuation of medications to avert an abstinence syndrome, treatment of preexistent pain, or preoperative initiation of therapy for postoperative pain management.
  • Anaesthetists offering perioperative analgesia services should provide, in collaboration with others as appropriate, patient and family education regarding their important roles in achieving comfort, reporting pain, and in proper use of the recommended analgesic methods.
  • Common misconceptions that overestimate the risk of adverse effects and addiction should be dispelled.
  • Patient education for optimal use of PCA and other sophisticated methods, such as patient-controlled epidural analgesia, might include discussion of these analgesic methods at the time of the preanaesthetic evaluation, brochures and videotapes to educate patients about therapeutic options, and discussion at the bedside during postoperative visits.
  • Such education may also include instruction in behavioral modalities for control of pain and anxiety.

4. Perioperative Techniques for Pain Management

  • Anaesthetists who manage perioperative pain should use therapeutic options such as epidural or intrathecal opioids, systemic opioid PCA, and regional techniques after thoughtfully considering the risks and benefits for the individual patient.
  • These modalities should be used in preference to intramuscular opioids ordered “as needed.”
  • The therapy selected should reflect the individual anesthesiologist’s expertise, as well as the capacity for safe application of the modality in each practice setting.
  • This capacity includes the ability to recognize and treat adverse effects that emerge after initiation of therapy.
  • Special caution should be taken when continuous infusion modalities are used because drug accumulation may contribute to adverse events.

5. Multimodal Analgesia

  • Whenever possible, anaesthetists should use multimodal pain management therapy.
  • Unless contraindicated, patients should receive an around-the-clock regimen of NSAIDs or paracetamol.
  • Regional blockade with local anaesthetics should be considered.
  • Dosing regimens should be administered to optimize efficacy while minimizing the risk of adverse events.
  • The choice of medication, dose, route, and duration of therapy should be individualized.

6. Patient Subpopulations

Paediatric patients

  • Aggressive and proactive pain management is necessary to overcome the historic under-treatment of pain in children.
  • Perioperative care for children undergoing painful procedures or surgery requires developmentally appropriate pain assessment and therapy.
  • Analgesic therapy should depend upon age, weight, and comorbidity, and unless contraindicated should involve a multimodal approach.
  • Behavioral techniques, especially important in addressing the emotional component of pain, should be applied whenever feasible.
  • Sedative, analgesic, and local anaesthetics are all important components of appropriate analgesic regimens for painful procedures.
  • Because many analgesic medications are synergistic with sedating agents, it is imperative that appropriate monitoring be used during the procedure and recovery.

Geriatric patients

  • Pain assessment and therapy should be integrated into the perioperative care of geriatric patients.
  • Pain assessment tools appropriate to a patient’s cognitive abilities should be used. Extensive and proactive evaluation and questioning may be necessary to overcome barriers that hinder communication regarding unrelieved pain.
  • Anaesthetists should recognize that geriatric patients may respond differently than younger patients to pain and analgesic medications, often because of comorbidity.
  • Vigilant dose titration is necessary to ensure adequate treatment while avoiding adverse effects such as somnolence in this vulnerable group, who are often taking other medications (including alternative and complementary agents).

Other subpopulations

  • Anaesthetists should recognize that patients who are critically ill, cognitively impaired, or have communication difficulties may require additional interventions to ensure optimal perioperative pain.management.
  • Anaesthetists should consider a therapeutic trial of an analgesic in patients with increased blood pressure and heart rate or agitated behavior when causes other than pain have been excluded.