You are on page 1of 21

101 MECHANICAL VENTILATION

NEIL R. MACINTYRE, MD

INTRODUCTION Positive-Pressure Ventilation and Cardiac Considerations in Choosing Ventilator


POSITIVE-PRESSURE MECHANICAL Function Settings for Different Forms of
VENTILATOR DESIGN FEATURES COMPLICATIONS OF POSITIVE- Respiratory Failure
Gas Delivery Systems PRESSURE MECHANICAL RECENT INNOVATIONS IN
Subsystems of Mechanical Ventilators VENTILATION MECHANICAL VENTILATORY
Ventilator-Induced Lung Injury SUPPORT
PHYSIOLOGIC EFFECTS OF POSITIVE-
PRESSURE MECHANICAL Oxygen Toxicity Innovative Strategies for “Lung
VENTILATION Patient-Ventilator Interface Complications Protection”
Ventilation and Respiratory System Patient-Ventilator Dys-synchrony Automated Weaning Strategies
Mechanics Pulmonary Infectious Complications Optimizing Synchrony During Interactive
Alveolar Recruitment and Gas Exchange Breaths
APPLYING MECHANICAL
Mechanical Loads VENTILATORY SUPPORT
Patient-Ventilator Interactions Mechanical Ventilatory Support Involves
Tradeoffs

INTRODUCTION tive with patient efforts. Pneumatic, electronic, or micro-


processor systems provide for various breath types. In
Mechanical ventilation is the process of using a device (ven- general, these can be classified by what initiates the breath
tilator) to support, partially or totally the delivery of gas to (trigger variable), what regulates gas delivery during the
the lungs. The desired effect of mechanical ventilation is to breath (target or limit variable), and what terminates the
maintain adequate levels of PO2 and PCO2 in arterial blood breath (cycle variable).5,6
while also unloading the inspiratory muscles. Although Breath triggers are generated by either a change in pres-
negative-pressure chambers or wraps might fulfill this defi- sure or flow initiated by patient effort (assisted/supported
nition, this discussion focuses on the use of devices that use breath) or by a set time (controlled breath). During the
positive airway pressures. breath, gas delivery is regulated to meet a target or limit
Positive-pressure mechanical ventilation is used widely. variable, which is generally either a set flow or a set inspira-
In the United States, about 1 to 3 million patients annually tory pressure. The breath is then terminated by a cycle vari-
are estimated to receive mechanical ventilatory support able, which can either be a set volume, a set inspiratory
outside the operating room.1 Traditionally, this support has time, or a set flow. A high-pressure cycle variable is usually
been provided in intensive care unit (ICU) settings, but there also present to limit lung over distention. Figure 101-1 uses
are clear trends toward expanding the venues to subacute this classification scheme to describe the five most common
facilities, long-term care facilities, and the home. As the breath types available on the current generation of ventila-
number of elderly increases and as more aggressive surgical tors: volume assist (VA), volume control (VC), pressure assist
and immunosuppressive therapies are developed, the need (PA), pressure control (PC), and pressure support (PS).
for mechanical ventilation in all of these venues is likely to
increase.1 In addition, increasing concerns about wide- Mode Controller/Feedback Systems
spread outbreaks of respiratory pandemics has led many The availability and delivery logic of different breath types
government agencies to stockpile large numbers of mechan- define the “mode” of mechanical ventilatory support.3,5,6
ical ventilators.2 The mode controller is an electronic, pneumatic, or
microprocessor-based system that is designed to provide the
proper combination of breaths according to set algorithms
POSITIVE-PRESSURE and feedback data (conditional variables) (Table 101-1).
The simplest mode is assist-control ventilation (ACV),
MECHANICAL VENTILATOR which can provide either flow-targeted volume-cycled
DESIGN FEATURES breaths (volume assist-control ventilation [VACV]) or pressure-
targeted time-cycled breaths (pressure assist-control ventila-
GAS DELIVERY SYSTEMS tion [PACV]). A simple feedback system is employed with
ACV that guarantees a set number of positive-pressure
Positive-Pressure Breath Controller breaths. If the patient’s underlying respiratory rate exceeds
Most modern ventilators utilize piston/bellows systems, tur- this guarantee, all breaths are patient-triggered breaths (VA
bines, or controllers of high-pressure sources to drive gas or PA breaths). If the patient’s respiratory rate is below this
flow.3,4 Tidal breaths are generated by this gas flow and can guarantee, the ventilator will “make up the difference” with
either be controlled entirely by the ventilator or be interac- mandatory (controlled) breaths (VC or PC breaths).
1761
1762 PART 3  •  Clinical Respiratory Medicine

VOLUME VOLUME PRESSURE PRESSURE PRESSURE


CONTROL ASSIST CONTROL ASSIST SUPPORT

Set pressure Set pressure Set pressure


pressure
Circuit

+
0

Set ti Set ti

Set flow Set flow Set minimal flow

in
Flow

0
out
increasing

Set volume Set volume


Volume

Machine Patient Machine Patient Patient


triggered triggered triggered triggered triggered
Figure 101-1  Circuit pressure, flow, and volume tracings over time depicting the five basic breaths available on most modern mechanical ventila-
tors. Breaths are classified by the variables that determine the trigger (machine time or patient effort), target/limit (ti, set flow or set pressure), and cycle
(set volume, set time, or set flow). The solid lines represent set or independent responses, and the dashed lines represent dependent responses. In all of
these breaths, pressure is usually a “backup” cycle variable designed to terminate gas delivery if circuit pressure rises above an alarm limit.

Table 101-1  Breath Types Available on Common Modes of apnea is detected; this represents a safety feature for modes
Mechanical Ventilation* with either low or no set guaranteed rates in case patient
BREATH TYPES AVAILABLE†
respiratory efforts are suddenly reduced or absent.
In recent years, more sophisticated feedback systems
Mode VC VA PC PA PS Sp have been developed for these basic modes and are now
Volume assist—control X X available on many modern devices. These include pressure-
Pressure assist—control X X regulated volume control (PRVC), volume support (VS), and
Volume SIMV X X X X adaptive support ventilation (ASV).
Pressure SIMV X X X X
PRVC (also known by proprietary names such as “VC+,”
“Autoflow,” and others) is an assist-control PACV that uses
Pressure support X
tidal volume as a feedback control for continuously adjust-
*In addition to the five “basic” breaths depicted in Figure 101-1, this table ing the pressure target.7 The clinician sets a tidal volume

also includes spontaneous unassisted/unsupported breaths. target, and the ventilator then automatically sets the inspi-
VC, volume control; VA, volume assist; PC, pressure control; PA, pressure ratory pressure within a clinician-set range to achieve
assist; PS, pressure support; Sp, spontaneous unassisted.
SIMV, synchronized intermittent mandatory ventilation.
this goal.
VS is also based on a feedback design using patient-
triggered, pressure-targeted, flow-cycled PS breaths. If a
patient’s respiratory drive exceeds the clinician-set guaran-
Another relatively simple mode is synchronized intermit- teed rate, some PRVC systems will provide additional
tent mandatory ventilation (SIMV), which can provide either patient-triggered, time-cycled PRVC breaths while others
flow-targeted volume-cycled breaths (volume SIMV) or will provide patient-triggered flow-cycled VS breaths. It is
pressure-targeted time-cycled breaths (pressure SIMV). Like important to note that, with both PRVC and VS breaths, an
ACV, SIMV guarantees a set minimal number of positive- improvement in respiratory mechanics will result in a lower
pressure breaths. Unlike ACV, however, if the patient’s respi- applied inspiratory pressure, whereas a worsening in respi-
ratory rate exceeds this guarantee, the ventilator will ratory mechanics will result in a higher applied inspiratory
provide assisted breaths up to the set rate and then allow pressure. Similarly, increasing respiratory effort by the
unassisted (simple SIMV mode) or flow-cycled pressure patient will result in a lower applied inspiratory pressure,
support breaths (SIMV + PS mode) thereafter. If the patient’s whereas decreasing respiratory effort by the patient will
respiratory rate is below the set guarantee, the ventilator result in a higher applied inspiratory pressure.
will again “make up the difference” with mandatory (con- ASV is also an assist-control, pressure-targeted, time-
trolled) breaths. Note that although PS breaths are often cycled mode of ventilation PACV that utilizes respiratory
provided during SIMV, PS breaths can also be provided as a system mechanics to set the tidal volume-frequency
stand-alone mode without any guaranteed breaths (pres- pattern.8 The clinician sets only a desired minute ventila-
sure support ventilation). Importantly, many modern tion and patient weight (for estimating anatomic dead
systems also have algorithms to generate breaths when an space). Using controlled breaths, ASV initially calculates
101  •  Mechanical Ventilation 1763

resistance and compliance, as well as the expiratory time open; this may result in inadvertent PEEP.14 As discussed in
constant (resistance × compliance). The ASV algorithm more detail later, a positive alveolar pressure may be present
then adjusts the frequency–tidal volume pattern to mini- at end expiration if the expiratory time is inadequate for
mize ventilator work (integral of pressure over volume) and the lung to return to its “resting volume” or if significant
thus conceptually to minimize applied forces to the lungs. flow limitation exists. Such inadvertent PEEP is referred to
The breathing pattern is also modulated by incorporating as intrinsic PEEP (PEEPi), or as “auto-PEEP,” “occult PEEP,”
the expiratory time constant to avoid air trapping. As respi- or “air trapping.”15
ratory mechanics change, the frequency–tidal volume
pattern is automatically adjusted to maintain minimal ven- Gas Delivery Circuit
tilatory work. In contrast, if patient efforts are triggering The circuit linking the ventilator to the patient usually con-
breaths with ASV, it behaves much like VS. sists of flexible tubing that often contains pressure or flow
sensors, closed suctioning systems, and an exhalation
valve. It is important to remember that, because this tubing
SUBSYSTEMS OF MECHANICAL VENTILATORS has measurable compliance (2 to 4 mL/cm H2O are repre-
sentative values), high circuit pressures may induce signifi-
Effort (Demand) Sensors cant amounts of delivered gas to distend the circuitry rather
Current ventilators have sensors that detect patient effort, than to enter the patient’s lungs.
thus allowing for a number of interactions between the
patient and the ventilator.9-11 Examples include patient- Patient–Ventilator Circuit Interface
triggered breaths in which the ventilator initiates flow in Positive-pressure ventilation is usually delivered through a
response to patient demand and pressure-targeted/limited tube inserted into the patient’s airway (orotracheal or naso-
breaths in which the ventilator adjusts flow in response tracheal tube or tracheostomy). These tubes generally have
to patient demand (see “Patient-Ventilator Interactions” air-filled balloons, which are inflated to provide a proper
later). These sensors are usually either pressure or flow airway seal. An alternative to the tracheal tube is a mask
transducers in the ventilatory circuitry and are character- system. Both full-face and nasal masks have been utilized
ized by their sensitivity (the circuit pressure or flow change with a variety of ventilatory support systems and modes.16
needed to initiate a ventilator response) and their responsive- Leaks around masks, however, can be significant and thus
ness (the time needed to provide this response).12 ventilatory support modes using masks must be able to
provide adequate volumes and proper inspiratory timing. To
Gas Blenders this end, special mask ventilators with pressure-targeted
Blenders mix air and O2 to produce a fractional concentration and either time-cycled or leak-compensated, flow-cycled
of O2 in inspired gas (FIO2) ranging from 0.21 to 1.0. On capabilities have been developed.16
newer systems, blenders are also available for other gases
such as heliox, nitric oxide (NO), and anesthetic agents. Aerosol Generators (see Chapter 11)
Therapeutic aerosols (e.g., bronchodilators, steroids, vaso-
Humidifiers dilators, antibiotics) can be delivered through the ventila-
With the upper airway bypassed by tracheal intubation, tor circuitry17 either by in-line nebulizers or by special
sufficient heat and moisture must be added to the inspired adapters designed for metered-dose inhalers. Lung deposi-
gas mixtures to avoid mucosal desiccation. Active humidi- tion is generally less in an intubated than in a nonintu-
fiers utilize external water sources and electrical power to bated patient because the endotracheal tube serves as a
adjust blended gas mixtures to near body conditions (tra- significant barrier to aerosol delivery. Higher dosing is thus
cheal temperature of > 35° C, water content of > 40 mg/L).13 advisable.
Heated wire circuits help facilitate this by preventing con- The location of the aerosol generator in the ventilator
densation and “rainout” in the ventilator tubing. Passive circuit can affect deposition. The optimal site appears to be
humidifiers use simple heat/moisture exchange devices in in the inspiratory limb several centimeters proximal to the
the ventilator circuit that reutilize heat and moisture patient’s “wye” connector.17 This location allows either
trapped from expired gas. These disposable units can usually nebulized medications or a metered-dose actuation to
supply adequate heat and moisture (i.e., > 30° to 33° C and “charge” the inspiratory limb of the circuit during exhala-
> 28 to 32 mg/L H2O) for many patients, particularly those tion. Aerosol particle velocity is also slowed and becomes
receiving mechanical ventilation for only short periods the leading portion of the next inspiration, both of which
of time.13 facilitate delivery.
Expiratory Pressure Generator Monitors and Graphic Displays
Positive end-expiratory pressure (PEEP), or positive airway Although electronic and microprocessor-based systems
pressure throughout expiration, can be generated to help have considerable internal monitoring of electronic and
maintain alveolar patency and improve ventilation-perfusion pneumatic function, the three variables generally displayed
  matching (see “Physiologic Effects of Positive-Pressure
( V/Q) for clinical use are circuit pressures, flows, and volumes.18
Mechanical Ventilation” later). PEEP is usually applied by Pressure sensors in the esophagus to estimate pleural pres-
regulating pressure in the expiratory valve of the ventilator sures are also available.19 Alarms can be utilized on all of
system but can also be applied by providing a continuous these monitors.6,20 Importantly, trending capabilities for up
flow of source gas during the expiratory phase. Some expira- to 72 hours or more are now available on many modern
tory valves have measurable resistance even when fully systems. Most modern positive-pressure ventilators also
1764 PART 3  •  Clinical Respiratory Medicine

have oxygen sensors in the circuitry to ensure that the Also, when V = 0 at end-inspiration, Pplat – PEEP allows
desired FIO2 is being delivered. In addition, some ventilators calculation of the static respiratory system compliance:
may also have analyzers for measuring exhaled carbon CRS = VT/(Pplat − PEEP)
dioxide and inhaled therapeutic gases such as NO or heliox.
Separating chest wall and lung compliance (CCW and CL,
respectively) requires measurement of esophageal pressure
PHYSIOLOGIC EFFECTS (Pes) to estimate pleural pressure.22a With this measure-
ment, the inspiratory change in Pes (ΔPes) can be used in
OF POSITIVE-PRESSURE the following calculations:
MECHANICAL VENTILATION CCW = VT/∆Pes
VENTILATION AND RESPIRATORY SYSTEM CL = VT/(Pplat − PEEP − ∆Pes)
MECHANICS (see Chapter 5) In clinical practice, because CCW is usually quite high and
ΔPes is thus quite low, Pplat alone is often taken as an
Alveolar Ventilation and the Equation of Motion approximation of end inspiratory lung-distending pressure.
Alveolar ventilation is the term for the delivery of fresh gas However, there are many situations in which there is a stiff
to the gas exchange regions of the lungs. Mathematically, chest wall and this crude approximation is not valid (e.g.,
this is expressed as: obesity, acute respiratory distress syndrome [ARDS], ascites,
surgical dressings). Under these situations, the impact of a
V A = f × (VT − VD)
stiff chest wall must be considered when using these mea-
where V A = alveolar ventilation, f = breathing frequency, surements to assess lung stretch.19,23,24
VT = tidal volume, and VD = wasted ventilation or dead space.
V A needs to be adequate to eliminate the carbon dioxide Flow-Targeted Versus Pressure-Targeted Breaths
production ( V CO2) while maintaining a reasonable arterial There are two basic approaches to delivering positive-
PCO2 (and pH) according to the following relationship: pressure breaths: flow targeting and pressure targeting (see
Fig. 101-1).3,5 With flow targeting (breaths 1 and 2 in Fig.
PaCO2 = (V CO2 /V A) × 800
101-1), the clinician sets the inspiratory flow; circuit pres-
The lungs are inflated by mechanical ventilation when pres- sure is then the dependent variable. With pressure targeting
sure and flow are applied at the airway opening. These (breaths 3 through 5 in Fig. 101-1), the clinician sets an
applied forces overcome respiratory system compliance inspiratory pressure target (with either time or flow as the
(both lung and chest wall components), airway resistance, cycling criterion); flow and volume are then the dependent
and respiratory system inertance and lung tissue resistance variables (i.e., varying with the lung mechanics and patient
to effect gas flow.21,22 For simplicity, because inertance and effort). With a flow-targeted breath, changes in compliance,
tissue resistance are relatively small, they can be ignored resistance, or patient effort will change Pcir (but not flow);
yielding the simplified equation of motion: in contrast, with a pressure-targeted breath, similar changes
in compliance, resistance, or effort will cause a change of
Driving pressure = (flow × resistance) tidal volume (but not Pcir) (see dashed lines in Fig. 101-1).
+ (volume/system compliance) Each strategy has advantages.25 For flow-targeted
In the mechanically ventilated patient, this relationship is breaths, a minimal tidal volume is guaranteed. For pressure-
expressed as: targeted breaths, the rapid initial flow and subsequent
∆Pcir + ∆Pmus = (V × R) + (VT/CRS) adjustable flow of pressure targeting may enhance gas
mixing and patient synchrony (see discussions “Distribu-
where ΔPcir is the change in ventilator circuit pressure tion of Ventilation” and “Patient-Ventilator Interactions”
above baseline (peak pressure minus set end-expiratory later).
pressure: Ppeak − PEEP); ΔPmus is patient inspiratory Pressure-targeted breaths can also be modified with a
muscle pressure generation (if present); V is the flow into volume feedback feature described earlier for PRVC and VS
the patient’s lungs; R is the resistance of the circuit, artifi- breaths to combine the enhanced gas mixing and patient-
cial airway, and natural airways combined; VT is the tidal ventilator synchrony effects of a pressure-targeted breath
volume; and CRS is the respiratory system compliance. If with a certain volume guarantee. However, it is important
there is PEEPi, it must be overcome by muscle and circuit to realize that providing a volume guarantee negates the
pressure before flow and volume can be delivered and thus pressure-limit feature set by the clinician because, with
PEEPi will add to the driving pressure requirement. PRVC/VS, worsening respiratory system mechanics will
During an inspiratory hold at end-inspiration in a patient drive applied pressures up. Another potential problem with
who is not making respiratory efforts (i.e., no-flow condi- PRVC/VS is that it may not adequately unload the patient
tion: V = 0, Pmus = 0), the ventilator circuit pressure “pla- if patient effort inappropriately increases due to pain or
teaus” at a pressure commonly referred to as the plateau anxiety.26
pressure (Pplat). In this way, the components of Pcir can be
determined. Specifically, the difference in Pcir during flow PEEPi and the Ventilatory Pattern
and during no-flow (the “peak to plateau difference”) allows PEEPi is the positive alveolar end-expiratory pressure that
for a calculation of total inspiratory resistance: arises because of inadequate expiration, caused by either
inadequate expiratory time or airway collapse during expi-
R = (Ppeak − Pplat)/V ration (or both). PEEPi increases with increased minute
101  •  Mechanical Ventilation 1765

NORMAL LOW REGIONAL CL NORMAL HIGH REGIONAL RAW

Positive Positive
pressure pressure
breath breath

A B
Figure 101-2  The distribution of ventilation in lung models with heterogeneous mechanical properties. Models of the lung are shown as two units
with homogeneous mechanical properties (normal): (A) with abnormal compliance distribution (low regional lung compliance [CL]) and (B) abnormal
resistance distribution (high regional airway resistance [RAW]). Note that in situations with heterogeneous lung mechanics, positive-pressure breaths are
preferentially distributed to “healthier” regions of the lung and can produce regional overdistention—even when a normal-sized tidal volume is delivered.
Note that in the obstructed example (high regional RAW), the overdistention may be transitory as gas moves from the low-resistance to the high-resistance
unit over time (pendelluft). (Redrawn from MacIntyre NR: Mechanical ventilatory support. In Dantzker D, MacIntyre NR, Bakow E, editors: Comprehensive respiratory
care. Philadelphia, 1995, WB Saunders, p 453.)

ventilation, decreased expiratory time fraction, and the compliance and low resistance and away from obstructed
increased respiratory system expiratory time constant (the or stiff units (Fig. 101-2A, Low Regional CL). This creates
product of resistance and compliance).27 the potential for regional overdistention of healthier lung
The development of PEEPi will have different effects on units, even in the face of “normal”-sized tidal volumes (see
flow-targeted compared with pressure-targeted ventilation. “Ventilator-Induced Lung Injury” later).
In flow-targeted ventilation, the constant delivered flow and The flow pattern set on the ventilator may also affect
volume (and thus ΔPcir) in the setting of a rising PEEPi will ventilation distribution. For example, when there are
increase both the Ppeak and the Pplat. In contrast, in marked inhomogeneities in airway resistance, slow and
pressure-targeted ventilation, the set Pcir limit coupled with constant flows will tend to distribute gas more evenly
a rising PEEPi level will decrease ΔPcir and thus the deliv- (although consequent shorter expiratory times may worsen
ered tidal volume (and minute ventilation). Importantly, air trapping).28 In addition, inspiratory pauses can also
this may help limit the build-up of PEEPi. allow pendelluft action to fill slowly filling alveoli (Fig.
In the patient without respiratory effort, PEEPi can be 101-2B, High Regional RAW). In contrast, when there is
recognized in two ways. First, when PEEPi is produced by parenchymal lung injury with minimal inhomogeneities in
an inadequate expiratory time, analysis of the flow graphic airway resistance, initially rapid flows with subsequent
will show that expiratory flow has not returned to zero deceleration (typically seen in pressure-targeted breaths)
before the next breath is given. Second, PEEPi in alveoli with may distribute gas more evenly and will pressurize lung
patent airways can be quantified during an expiratory hold units rapidly, producing a higher mean inspiratory alveolar
maneuver that permits equilibration of the PEEPi with pressure for a given breath volume.30
Pcir.27 It should be noted that more uniform ventilation distribu-
In the patient with active respiratory effort, PEEPi can be tion does not necessarily mean better V/Q  matching (i.e.,
assumed to be present if the expiratory flow has not reached more homogeneous ventilation distribution may actually
zero before the next breath begins. However, the expiratory worsen V/Q  matching in a lung with inhomogeneous per-
hold maneuver cannot be interpreted in an actively breath- fusion). Because of all these considerations, predicting
ing patient. As noted in more detail later, in an actively which flow pattern will optimize ventilation-perfusion
breathing patient, PEEPi can function as an inspiratory matching is difficult and often requires trial and error.
threshold load. This is best quantified by using Pes to esti-
mate pleural pressures. With this technique, the change in ALVEOLAR RECRUITMENT AND GAS EXCHANGE
Pes before a change in Pcir is a reflection of the threshold
load imposed by PEEPi (see “Patient-Ventilator Interac- Because of alveolar flooding, inflammatory exudates,
tions” later). and collapse, parenchymal lung injury leads to V/Q   mis-
matching and shunts.31 In many (but not all) of these
Distribution of Ventilation disease processes, substantial numbers of collapsed alveoli
A positive-pressure tidal breath must be distributed among can be recruited during a positive-pressure ventilatory
the millions of alveolar units in the lung.28,29 Factors affect- cycle.32-35Additional recruitment can sometimes be pro-
ing this distribution include patient-related factors: regional vided with the use of formal recruitment maneuvers or pro-
resistances, compliances, and functional residual capaci- longation of inspiratory time.36,37 The application of PEEP is
ties. Breaths will tend to distribute more to units with high designed to prevent derecruitment during exhalation.
1766 PART 3  •  Clinical Respiratory Medicine

recruitment is lower than that required for initial


Recruitment Maneuvers recruitment.
Recruitment maneuvers (RMs) can be performed using sus- PEEP, however, can also be detrimental. Because the tidal
tained inflations (e.g., 30 to 40 cm H2O) for up to 30 to 120 breath is delivered on top of the baseline PEEP, end-
seconds, by transient elevations of the PEEP-tidal volume inspiratory pressures are usually raised by the application
settings, and by single or multiple “sighs” that take the lung of PEEP (although this increase may be less than the actual
briefly to near total lung capacity.36 To avoid patient inspira- added PEEP because of improved compliance). This increase
tory or expiratory efforts, additional sedation or even neu- must be considered if the lung is at risk for regional over-
romuscular blockade may be used. Importantly, RMs can distention (see “Ventilator-Induced Lung Injury” later).
have adverse hemodynamic consequences; close monitor- Moreover, since parenchymal lung injury is often quite het-
ing of the patient is mandatory during RMs. RMs provide erogeneous, appropriate PEEP in one region may be subop-
initial alveolar recruitment only—maintenance of recruit- timal in another and excessive in yet another.35,43,44
ment requires setting PEEP appropriately to prevent subse- Optimizing PEEP is thus a balance between recruiting the
quent derecruitment.36 recruitable alveoli in diseased regions without overdistend-
ing already recruited alveoli in healthier regions. Another
Inspiratory Time Prolongations potential detrimental effect of PEEP is that it raises mean
A positive-pressure breath produces a flow magnitude and intrathoracic pressure, thereby compromising cardiac
a flow profile that, as noted previously, can affect ventilation filling in susceptible patients (see “Positive-Pressure Ventila-
  Prolonging inspiratory time,
distribution (and thus V/Q). tion and Cardiac Function” section later).
generally by adding a pause, often in conjunction with a
rapid decelerating flow (i.e., pressure-targeted breath), has
MECHANICAL LOADS
several physiologic effects. First, the longer inflation period
may recruit more alveoli.38,39 Second, increased gas mixing Mechanical loads describe the physical requirements of
time may improve V/Q   matching in parenchymal lung ventilation with a single value, either the pressure-time
38
injury (pendelufft). Third, the development of PEEPi from product (PTP—the integral of pressure over time) or work
consequently shorter expiratory times can have effects (W—the integral of pressure over volume).21,22 Because
similar to those of applied PEEP (see earlier).39 It should be mechanical loads correlate with inspiratory muscle oxygen
noted, however, that the distribution of PEEPi, which is demands,45-47a the concept of load is useful in considering
most pronounced in lung units with long expiratory time inspiratory muscle energy requirements during spontane-
constants, may be different from that of applied PEEP and ous or interactive ventilatory support. Moreover, as
  may also be different with intrinsic
thus the effects on V/Q described in more detail later, load referenced to muscle
compared with applied PEEP. Fourth, because these long strength and/or endurance properties (e.g., PTP or W
inspiratory times significantly increase total intrathoracic divided by muscle pressure-generating capability) can be
pressures, cardiac output may be reduced (see “Positive- used to set levels of partial ventilatory support or predict
Pressure Ventilation and Cardiac Function” later). And spontaneous breathing capabilities.47
finally, inspiratory-to-expiratory ratios that exceed 1 : 1 Compliance, resistance, flow, and volume all contribute
(so-called inverse-ratio ventilation) are uncomfortable, to the magnitude of the load per breath. During spontane-
and patient sedation/paralysis is often required unless a ous breaths, Pcir is zero and integrating Pes over time or
relief mechanism allows spontaneous breathing during the volume (referenced to the passive inflation pressure)
inflation period (see “Airway Pressure Release Ventilation” describes the load borne by the inspiratory muscles to
later). inflate the lungs. During a controlled breath, integrating
Pcir over time or volume describes the load borne by the
Positive End-Expiratory Pressure ventilator to inflate the entire respiratory system (lungs
PEEP is defined as an elevated airway pressure at the end and chest wall) and integrating Pes over time or volume
of expiration.32 As noted earlier, PEEP can be produced describes the loads imposed by the chest wall only. During
by either expiratory circuit valves (applied PEEP) or inade- interactive breaths the load is shared between patient and
quate expiratory times in lung units with long expiratory ventilator.48
time constants (PEEPi).21,22,38 Note that expiratory muscle Under heavy loading conditions (e.g., the patient with
contraction can also raise intrathoracic pressures at end- abnormal respiratory system mechanics and thus high
expiration, but this does not have the same effects on the pressure requirements), the duration of pressure (i.e., the
lungs because transpulmonary pressure is not increased. PTP) correlates better with muscle energetics and fatigue
PEEP helps to recruit or maintain alveolar units open, potential than does the volume moved with pressure (i.e.,
providing several potential benefits. First, recruited alveoli work).45,46 Indeed, during ventilation requiring high pres-
  matching and gas exchange throughout the
improve V/Q sures, multiplying PTP by the inspiratory time fraction and
ventilatory cycle.32 Second, as discussed in more detail later, referencing this to the maximal pressure that the inspira-
patent alveoli throughout the ventilatory cycle are not tory muscles can generate results in the pressure-time index
exposed to the risk of injury from the shear stress of repeated (PTI). Muscle fatigue can be expected if the PTI value
opening and closing.40,41 Third, open alveoli with intact sur- exceeds 0.15.48,49 The concern with high pressure loads in
factant monolayers improve lung compliance.42 This is the patients receiving partial ventilatory support is one of the
rationale behind applying PEEP after an RM: recruited rationales for providing ventilator pressure assistance with
alveoli are on the deflation limb of the pressure-volume every spontaneous effort (i.e., pressure-assisted or sup-
relationship and thus the pressure required to maintain ported breaths), as opposed to supporting only some breaths
101  •  Mechanical Ventilation 1767

as in intermittent mandatory ventilation without pressure sion pressures in the three-zone West lung model helps
support (see “Patient-Ventilator Interactions” later).50 explain this.61 Specifically, the dependent lung is generally
Inspiratory muscle overload is one of the major determi- in a zone 3 (capillary distention) state. As intra-alveolar
nants of continuing ventilator dependency and can result pressures rise, however, zone 2 and zone 1 (capillary
either from excessive mechanical loads or from inspiratory  
collapse/dead space) regions can appear, creating high V/Q
muscle dysfunction. Excessive mechanical loads can result units. Indeed, increases in dead space can be a consequence
from disease or from inappropriate ventilatory assistance of ventilatory strategies employing high ventilatory pres-
(see “Patient-Ventilator Interactions” later). Clinically, sures, as well as with those producing PEEPi.
inspiratory muscle overload is manifested by rapid, shallow Positive-pressure mechanical ventilation can affect other
breathing patterns, paradoxical abdominal motion, and aspects of cardiovascular function. Specifically, dyspnea,
patient distress. Inspiratory muscle dysfunction can be a anxiety, and discomfort from inadequate ventilatory support
result of the systemic inflammatory response syndrome, can lead to stress-related catecholamine release with subse-
metabolic disturbances, drugs (e.g., steroids, previous use quent increases in myocardial oxygen demands and risk of
of neuromuscular blockers), malnutrition, or malposition- dysrhythmias.60 In addition, coronary blood vessel oxygen
ing (i.e., diaphragm flattening from lung overinflation).51 delivery can be compromised by inadequate gas exchange
Finally, insufficient loading may also affect inspiratory from the lung injury coupled with low mixed venous PO2
muscles. Specifically, controlled mechanical ventilation due to high oxygen consumption demands by the inspira-
without any patient effort, perhaps for as little as 24 hours, tory muscles.
may produce muscle changes similar to disuse atrophy—a
condition described as ventilator-induced diaphragmatic dys-
function (VIDD).52-54
COMPLICATIONS OF
PATIENT-VENTILATOR INTERACTIONS POSITIVE-PRESSURE
Mechanical ventilation modes that permit spontaneous MECHANICAL VENTILATION
ventilatory activity are termed “interactive” modes. These
interactions can range from simple triggering of mechani- VENTILATOR-INDUCED LUNG INJURY
cal breaths to more complex processes affecting delivered The lung can be injured when it is stretched excessively.
flow patterns and breath timing. Interactive modes allow for The most obvious injury is barotrauma: alveolar rupture
muscle “exercise,” which, when performed at nonfatiguing presenting as extra-alveolar air in the mediastinum (pneu-
or physiologic levels, may prevent VIDD and facilitate momediastinum), pericardium (pneumopericardium), sub-
fatigue recovery.11,52-54 In addition, permitting spontaneous cutaneous tissue (subcutaneous emphysema), pleural
patient ventilatory activity with “comfortable” interactive space (pneumothorax), or vasculature (air emboli).62 The
modes may reduce the need for the sedation and/or neuro- risk for extra-alveolar air increases as a function of the
muscular blockers that may otherwise be used to prevent magnitude and duration of alveolar overdistention. Thus
patients from “fighting the ventilator.”11,55 Interactions take interactions of respiratory system mechanics and mechan-
place during all three phases of breath delivery: breath trig- ical ventilation strategies (high regional VT and PEEP—
gering, flow delivery, and breath cycling and are described both applied and intrinsic) that produce regions of
in detail later in “Patient-Ventilator Dys-synchrony.” excessive alveolar stretch create alveolar units at risk for
rupture.
POSITIVE-PRESSURE VENTILATION AND Even without producing extra-alveolar air and rupture,
mechanical ventilation can induce ventilator-induced lung
CARDIAC FUNCTION
injury (VILI).62a,62b In experimental animals, acute lung
In addition to affecting ventilation and ventilation distribu- injury can be produced by mechanical ventilation strategies
tion, positive-pressure ventilation can also affect cardio­ that stretch the lungs beyond the normal maximum volume
vascular function.56-58 In general, as mean intrathoracic (at transpulmonary distending pressures of 30 to 35 cm
pressure increases, right ventricular filling decreases H2O).63-65 In engineering parlance, this is termed mechani-
and cardiac output/pulmonary perfusion consequently cal “stress.”66 A number of clinical trials clearly indicate
decreases. This is the rationale for using volume repletion that ventilator strategies exposing the injured human lung
to maintain cardiac output in the setting of high intratho- to transpulmonary pressures in excess of 30 to 35 cm H2O
racic pressure. Of note, the effect of reduced cardiac filling are associated with lung injury.67-70 Importantly, VILI may
on cardiac output may be partially counteracted by improve- be more than simply a consequence of excessive end-
ment in left ventricular function due to elevated intratho- inspiratory stretch. Even in the setting of transpulmonary
racic pressure, which can reduce left ventricular afterload.59 pressures less than 30 cm H2O, excessive tidal stretch from
Importantly, in patients with left heart failure, the reduced repetitive cycling of the lung with tidal volumes in excess of
cardiac filling and reduced left ventricular afterload effects 8 to 10 mL/kg ideal body weight (IBW) may contribute to
of elevated intrathoracic pressure may improve cardiac VILI.67,68 Interestingly, this VILI risk may be better quanti-
function to the point that the removal of intrathoracic pres- fied by referencing the tidal volume to the resting lung
sure may worsen cardiac function and thereby produce volume and can be termed mechanical “strain.”66 Other
weaning failure.60 ventilatory pattern factors, such as the frequency of
Intrathoracic pressures can also influence distribution of stretch71 or the acceleration/velocity of stretch,72 may also
perfusion. The relationship of alveolar pressures to perfu- be involved in the development of VILI. VILI appears to be
1768 PART 3  •  Clinical Respiratory Medicine

potentiated by a shear stress phenomenon that takes place VILI is manifest pathologically as diffuse alveolar
when injured alveoli repetitively open and collapse during damage.40,63,64,79 Moreover, VILI is associated with cytokine
the ventilatory cycle (i.e., cyclical atelectasis).40,73,74 Vascu- release79,80 and bacterial translocation,81 which are impli-
lar pressure elevations may also contribute to VILI.75 cated in the systemic inflammatory response with multior-
VILI likely develops regionally when low-resistance/ gan dysfunction that results in VILI-associated mortality.
high-compliance units receive a disproportionately high
regional tidal volume in the setting of high alveolar distend- OXYGEN TOXICITY
ing pressures (see Fig. 101-2). This can be appreciated
using computed tomography under circumstances when Very high inspired oxygen concentrations can cause oxidant
pressures required for recruitment of diseased atelectatic injury in airways and lung parenchyma.82 Much of the data
regions simultaneously produce overdistention in already supporting this concept, however, has come from animals
open, less-diseased regions (Fig. 101-3).76 Regional protec- that often have quite different tolerances to oxygen than
tion of these healthier lung units is the rationale for using humans. It is thus not clear what the “safe” oxygen concen-
“lung-protective” ventilator strategies that accept less than tration or duration of exposure is in sick humans. Most
normal values for pH and arterial PCO2 in exchange for consensus groups have argued that FIO2 values less than
lower (and safer) distending pressures (see “Applying 0.4 are safe for prolonged periods of time and that FIO2
Mechanical Ventilatory Support” later).77 Interestingly, values of greater than 0.70 should be avoided if possible.
data suggest that a permissive respiratory acidosis might Interestingly, several observational trials suggest that even
also have therapeutic effects on alveolar injury from VILI, with a “safe” FIO2 less than 0.4, maintaining an arterial PO2
although the clinical applicability of this observation is greater than 120 to 130 mm Hg may produce systemic
not clear.78 oxygen toxicity over time.83,84

1500 450

400
1250 high PEEP
350
Lung volume (mL)

1000 300

250
750 moderate PEEP
200

500 150
no PEEP
100
250
50

0 0
0 5 10 15 20 25 30 −900 −700 −500 −300 −100 0 100
Airway pressure (cm H 2 O) Hounsfield units (HU)

moderate
no PEEP PEEP high PEEP

Figure 101-3  Effects of positive airway pressure in a lung with heterogeneous injury. The upper left panel shows the almost linear pressure-volume
relationships as positive airway pressure is applied to the lungs. At the bottom are three representative CT lung slices at low, midrange, and maximal
airway pressure. In the lung at low airway pressure (no PEEP, lower left), note the substantial atelectasis in the dependent lung regions, which is progres-
sively reduced as airway pressure is increased (moderate PEEP and high PEEP, in the right two images). The upper right panel depicts the distribution of
Hounsfield units (HUs) in each of these three CT slices, at low pressure (dark blue), at midrange airway pressure (brown), and at the highest airway pressure
(light green). Note that, although increasing pressure progressively reduces the number of atelectatic units (those with CT numbers near zero), these same
pressures simultaneously increase the number of overdistended lung units elsewhere in the lungs (those with CT numbers < −900 HU). (From Vieira SR,
Puybasset L, Lu Q, et al: A scanographic assessment of pulmonary morphology in acute lung injury. Am J Resp Crit Care Med 159:1612–1623, 1999.)
101  •  Mechanical Ventilation 1769

PATIENT-VENTILATOR INTERFACE mean that muscle loads are eliminated. Instead, it means
COMPLICATIONS that a delivered flow pattern is associated with a muscle
loading pattern that resembles normal, comfortable breath-
The patient must be connected to the ventilator via ventila- ing11 (“synchronous” flow delivery).
tor circuitry and an artificial airway. Problems with this In general, flow synchrony is best assessed by clinical
interface can lead to complications. The most obvious issue assessment and analyzing the graphic representation of
is disconnection from the ventilator (including artificial circuit pressure over time.11,92 Flow synchrony is manifest
airway dislodgment). Disconnections have been reported in clinically as a relaxed patient who does not appear dyspneic.
up to 8% to 13% of ventilated patients85 and, if left uncor- In contrast, flow dys-synchrony is often manifest as tachy-
rected, can be fatal. Because circuit pressure and flow can pnea, air hunger (dyspnea), and the appearance of “flow
be maintained despite the disconnection of the ventilator starvation.”11 With synchronous flow delivery, the circuit
from the patient (e.g., if the airway is in the esophagus or if pressure graph should maintain either a steady baseline
the disconnected circuit remains partially occluded), it is profile (a CPAP breath) or a convex upward shape (assisted/
critical that carefully set, redundant (i.e., pressure, flow, and supported breath) indicating that flow is proportional to
even exhaled carbon dioxide) alarms are present.6 Other demand. Dys-synchrony (and imposed loading) is said to
complications of the patient-ventilator interface include exist when the circuit pressure graph is literally “sucked
obstructions from secretions, circuit leaks, airway injury downward” by effort in excess of flow delivery, often below
from inadequate heat/humidity, tracheal injury from the the baseline11,93 (Fig. 101-4).
artificial airway, and loss of delivered tidal volume in a com-
pliant circuit. Breath Cycling
Cycling dys-synchrony can arise in one of two ways. First,
if the mechanical breath lasts beyond the duration of
PATIENT-VENTILATOR DYS-SYNCHRONY
patient effort, an inadequate expiratory time may develop
Patient-ventilator dys-synchrony describes the delivery of a (along with air trapping) and/or patient expiratory efforts
breath from a mechanical ventilator that is not matched to may be required to terminate the breath.11 This can be a
patient effort. As noted earlier, this can take place during particular concern when using PS breaths in obstructive
the triggering process, flow delivery, and breath cycling. airway disease—the relatively steady inspiratory flow with
PS in these patients coupled with the PS flow-cycling algo-
Ventilator Breath Triggering rithm may significantly delay breath termination.94 Second,
Ventilators may sense a spontaneous effort by either an if the mechanical breath terminates before the patient
airway pressure drop or an airway flow change.9,10,86 Even effort is finished, the patient may be left demanding addi-
with modern sensors, there is unavoidable dys-synchrony tional flow without any being delivered. Significant imposed
in the triggering process. First, a certain level of sensor
insensitivity must be incorporated to avoid artifacts trigger-
ing the ventilator (i.e., “auto-triggering” due to cardiogenic
Dys-synchronous Synchronous
oscillations). Second, even when the patient effort has been
sensed, there is an inherent delay (up to 100 msec or more) · ·
V V
in activating a valving system to open and achieve target
airway flow (system responsiveness). Both of these factors
Flow
can result in significant “isometric-like” pressure loads on
the inspiratory muscles during the triggering process. In
addition, in the setting of air trapping and PEEPi, the ele- VT VT
vated end-expiratory alveolar pressure acts as a triggering
threshold load on the inspiratory muscles. Under these con- Volume
ditions, judicious use of applied PEEP can equilibrate expi-
ratory pressure throughout the lungs and the ventilator PAW PAW
circuit to reduce this triggering load.87,88
Excessive triggering can also result from auto-triggering
Pressure
as noted earlier or from the triggering of a second breath
when the mechanical breath cycles before the termination 2 4 6 2 4 6
of patient effort (“breath stacking”). A recently described
phenomenon, “entrainment,” can also result in a double Figure 101-4  Flow and pressure pattern differences during dys-
breath.89 Entrainment can be seen during a machine- synchronous and synchronous breaths in a patient with a vigorous
triggered breath when the delivered gas flow elicits an effort. inspiratory effort. Depicted are flow (upper panel), volume (middle panel),
Sometimes this simply extends the breath, but if the effort and pressure (lower panel) for a dys-synchronous flow-targeted breath
(left) and a more synchronous pressure-targeted breath (right) with
persists beyond machine breath termination, a second matched mean flow, inspiratory time, and tidal volume. Note that, during
breath can be triggered. the flow-targeted breaths (left), the fixed flow does not respond to the
vigorous inspiratory effort, resulting in the circuit pressure being “sucked”
Ventilator-Delivered Flow Pattern downward (solid arrow). In contrast, with the pressure-targeted breath,
flow adjusts and increases to meet the inspiratory effort better (dashed
During an interactive breath, inspiratory muscles are con- arrow). (Redrawn from Yang LY, Huang YC, MacIntyre NR: Patient-ventilator
tracting90,91 and the ventilator flow delivery should be ade- synchrony during pressure-targeted versus flow-targeted small tidal volume
quate to provide proper muscle unloading. This does not assisted ventilation. J Crit Care 22:252–257, 2007.)
1770 PART 3  •  Clinical Respiratory Medicine

loading and, as noted earlier, double breath triggering may modes that avoid multiple different breath types (i.e., avoid-
result.11 ing SIMV) tend to facilitate flow synchrony.11 Finally, breath
duration should be optimized for comfort and for elimina-
Clinical Implications tion of double breaths by using adjustments in the cycle
Determining the prevalence of patient-ventilatory dys- variable (volume, time, or flow).
synchrony is difficult because studies examining this Two new modes (proportional assist ventilation [PAV] and
question have used heterogeneous patient populations neurally adjusted ventilatory assistance [NAVA]) have been
and different definitions of dys-synchrony, methods of used over the past decade to enhance synchrony further.
detection, duration and timing of observation, and ventila- These are described in more detail later in “Recent Innova-
tory modes.11,95 Triggering dys-synchronies have been the tions in Mechanical Ventilatory Support.”
most well studied. Depending on patient population, venti-
lator settings, and measurement techniques, triggering PULMONARY INFECTIOUS COMPLICATIONS
dys-synchronies have been reported in 26% to 82% of
mechanically ventilated patients. Not surprisingly, trigger Mechanically ventilated patients are at risk for pulmonary
dys-synchronies were more common in patients with COPD infections for several reasons.97 First, the natural glottic
and those at risk for intrinsic PEEP development.11,95 Double closure protective mechanism is compromised by an endo-
triggering is the other commonly reported triggering dys- tracheal tube. This permits continuous seepage of oropha-
synchrony but is generally described in less than 10% of ryngeal material into the airways. Second, the endotracheal
patients in these various studies.11,95 tube itself impairs the cough reflex and serves as an addi-
The incidence of other forms of dys-synchrony (flow dys- tional potential portal for pathogens to enter the lungs. This
synchrony and cycle dys-synchrony) has not been as well is particularly important if the circuit is contaminated.
characterized. However, a retrospective evaluation of the Third, airway and parenchymal injury both from the under-
National Institutes of Health ARDS Network small tidal lying disease and from management complications make
volume study reported cycling dys-synchronies associated the lung prone to infections. Fourth, the ICU environment
with double triggering in 9.7% of all breaths analyzed.96 itself, with its heavy antibiotic use and presence of sick
Indeed, it is likely that patient-ventilatory dys-synchrony is patients in close proximity, increases risk for a variety of
ubiquitous if any patient is observed long enough during infections.
assisted/supported mechanical ventilation, especially if Preventing ventilator-associated pneumonia (VAP) is criti-
more sophisticated monitors (e.g., Pes or diaphragmatic cally important because the development of VAP heavily
electromyograms) are employed to detect patient effort. influences both length of stay and mortality.97-100 Care
Although there is no doubt that many dys-synchronies “bundles” linked to better outcomes include handwashing,
are subtle and of little clinical relevance, significant dys- elevating the head of the bed, oral care with chlorhexidine,
synchronies may be missed, thus overloading respiratory and carefully chosen antibiotic regimens for other infec-
muscles and producing patient discomfort, which is a tions. Management strategies that avoid breaking the integ-
frequently cited indication for the administration of seda- rity of the circuit (e.g., by changing the circuit only when
tives.11,95 This may impact ventilator duration because high visibly contaminated) also appear to be helpful.98,99 Con-
sedation usage is linked to longer ventilator use. Indeed, tinuous drainage of subglottic secretions may be another
several observational studies have noted more days of simple way of reducing lung contamination by oropharyn-
mechanical ventilation and even trends toward higher mor- geal material.98,99 More controversial are devices to cleanse
tality in patients with trigger dys-synchronies taking place the endotracheal tube or make it resistant to biofilm
more than 10% of the time.95 buildup.101
Another concept explored in small trials is the use
Managing Dys-synchronies of aerosolized antibiotics in patients with purulent secre-
Synchronous ventilator settings first require setting the tions to reduce progression of tracheobronchitis to
trigger to as sensitive a setting as possible (without auto- ventilator-associated pneumonia.102 Finally, prompt discon-
triggering). If PEEPi is creating a triggering load, attempts tinuation of ventilatory support when clinically appropri-
should be made to minimize PEEPi and then judicious use ate will help minimize the time patients are exposed to
of applied PEEP added to counterbalance the imposed trig- infection risks.
gering load as described later. If entrainment is present,
reducing sedation and the need for a mandatory (con-
trolled) breath rate should be re-assessed. APPLYING MECHANICAL
Flow synchrony is often easier to achieve with variable
flow pressure-targeted breaths, as noted earlier. Pressure-
VENTILATORY SUPPORT
targeted breaths can also allow for adjustments in the rate MECHANICAL VENTILATORY SUPPORT
of pressure rise and can compensate for endotracheal tube
resistance, as means of further enhancing synchrony (see INVOLVES TRADEOFFS
“Recent Innovations in Mechanical Ventilatory Support” The goals of providing adequate support while minimizing
later). If flow-targeted breaths are desired, synchrony can the risk of VILI and other complications involves tradeoffs.
be addressed with adjustments in flow magnitude and Specifically, the need for potentially injurious pressures,
profile (sine wave, square wave, decelerating pattern). volumes, and supplemental O2 must be weighed against the
Importantly, because the patient’s respiratory drive is mod- benefits of supporting gas exchange. To this end, a
ulated by mechanical feedback from the lungs and thorax, “re-thinking” of gas exchange goals has taken place over
101  •  Mechanical Ventilation 1771

the past 2 decades so that now, pH levels as low as 7.15 to diseases, the potential for air trapping in parenchymal lung
7.20 and PO2 values as low as 55 mm Hg are often consid- injury is low if the breathing frequency is less than 35
ered acceptable in order to protect the lung.70,78,103 Ventila- breaths/min; air trapping may not develop even at frequen-
tor settings are thus selected to provide at least this level of cies exceeding 50 breaths/min.
gas-exchange support while meeting three mechanical There is controversy as to whether neuromuscular blockade
goals: (1) provision of enough PEEP to recruit the “recruit- (NMB) and totally machine-controlled mechanical ventila-
able” alveoli, (2) avoidance of a PEEP-tidal volume combi- tion should be used in the initial 24 to 48 hours of paren-
nation that unnecessarily overdistends lung regions at chymal lung injury. NMB use producing flaccid ventilatory
end-inspiration, and (3) limiting tidal volumes to the physi- muscles reduces total body oxygen demands and eliminates
ologic range. These goals embody the concept of a lung- potential dys-synchronous interactions. Indeed, one study
protective mechanical ventilatory strategy. Currently, these in patients with severe hypoxemia (arterial PO2/FIO2 ratios
principles guide recommendations for the specific manage- < 120) showed improved mortality with NMB use for 48
ment of various forms of respiratory failure.104-107 hours.110 However, as noted earlier, ventilatory muscles
inactive for as little as 24 hours are at risk for VIDD and
long-term disability.52-54 Moreover, the use of NMB usually
CONSIDERATIONS IN CHOOSING VENTILATOR requires substantial sedative use. As a consequence, many
SETTINGS FOR DIFFERENT FORMS OF authorities would argue that assisted/supported breaths
RESPIRATORY FAILURE are preferable early in the course of mechanical ventilatory
support. The inspiratory time and the I : E ratio in parenchy-
Parenchymal Lung Injury mal injury is set with consideration of several issues. The
Parenchymal lung injury means injury that involves the air usual initial I : E ratio setting is 1 : 2 to 1 : 4, the normal and
spaces and the interstitium of the lung.31-35 In general, the most comfortable setting. The flow graphic should also
parenchymal injury stiffens lungs and decreases lung be assessed to ensure that cycle synchrony is present and
volumes. It is important to realize that there are often the expiratory time is adequate to avoid air trapping. An I : E
marked regional differences in the degree of mechanical ratio greater than 1 : 1 is referred to as inverse-ratio ventila-
abnormalities, which interact with the particular ventila- tion (IRV). In severe respiratory failure, IRV can be employed
tory strategy. This is because delivered gas will preferentially as an alternative to increasing PEEP to improve V/Q   match-
go to the more normal regions, those with higher compli- ing.38,39 The beneficial mechanisms involved include longer
ance and lower resistance, rather than to abnormal regions mixing times, recruitment of slowly filling alveoli, and
(see Fig. 101-2). A “normal-sized” tidal volume may thus development of PEEPi. A variation on IRV is airway pressure
be distributed preferentially to the healthier regions, result- release ventilation (APRV).111-115 APRV incorporates the
ing in a regional overdistention injury. Parenchymal injury ability to breathe spontaneously during the long inflation
can also affect the airways, especially the bronchioles and period of a pressure-controlled breath—a feature that may
alveolar ducts. These narrowed and collapsible small enhance recruitment and gas mixing. APRV is discussed in
airways can also reduce regional ventilation to injured lung more detail in “Recent Innovations in Mechanical Ventila-
units, leading to regions of air trapping and possibly cyst tion” later.
formation during the healing phase. The PEEP/FIO2 settings are optimized using both mechan-
Gas-exchange abnormalities in parenchymal lung injury ical and gas-exchange considerations. Conceptually, in
are a consequence of alveolar consolidation, flooding, and/ setting the PEEP level in parenchymal lung injury, the goal
or collapse producing a maldistribution of ventilation and is to provide ventilator settings between the upper and lower
resulting in V/Q  mismatching and shunts.31-35 Because inflection points on the deflation limb of the pressure-
  mismatching and shunt are more of an issue than
V/Q volume curve.116 The most direct mechanical approaches
dead space in parenchymal lung disease, hypoxemia tends use the static pressure-volume plot to set the PEEP and VT.
to be more of a clinical problem than carbon dioxide This traditionally involves multiple VT/Pplat measurements
elevation. and requires considerable clinician time along with patient
Frequency-tidal volume settings for patients with parenchy- sedation or even neuromuscular blockade and is not widely
mal lung injury must focus on limiting end-inspiratory used. Another mechanical approach uses step changes in
stretch. The benefit of reducing stretch in improving PEEP to determine the PEEP level that gives the best compli-
outcome has been suggested by several clinical trials67-70,108 ance.117 A simpler mechanical approach analyzes the Pcir
but has been most convincingly demonstrated by the profile during a constant-flow breath (stress index) to detect
National Institutes of Health (NIH)-sponsored ARDS Network overdistention (late rising profile) or collapse/reopening
Trial in which a ventilator strategy using a VT of 6 mL/kg injury (early rising profile).118 With all of these approaches,
predicted body weight (PBW) as compared with 12 mL/kg a recruitment maneuver can be used to recruit the maximal
PBW led to a 10% absolute reduction in mortality.70 Thus number of recruitable alveoli before setting the PEEP. FIO2
initial VT settings should start at 6 mL/kg PBW.68,109 More- adjustments are then set as low as clinically acceptable.
over, strong consideration should be given to reducing this PEEP can also be guided by gas-exchange criteria, usually
setting further if end-inspiratory Pplat exceeds 30 cm involving algorithms that adjust PEEP and FIO2 according
H2O.67-70 Increases in the VT settings might be considered if to certain targets. Note that constructing a PEEP/FIO2
there is marked patient discomfort or suboptimal gas algorithm is usually an empirical exercise in balancing dis-
exchange provided that the subsequent Pplat values do not tending pressure, arterial O2 saturation, and FIO2 and
exceed 30 cm H2O.67-70 Respiratory rate settings are then depends on the clinician’s perception of the relative “toxici-
adjusted to control pH. Unlike the case in obstructive ties” of high thoracic pressures, high FIO2, and low arterial
1772 PART 3  •  Clinical Respiratory Medicine

24 tial mechanical disadvantage that further worsens muscle


function. Overinflated regions may also compress more
PEEP (cm H2O)
20
16 healthy regions of the lung, impairing V/Q   matching.
12 Regions of air trapping and PEEPi also function to increase
8 the threshold load the patient must overcome to trigger
4 mechanical breaths, as noted earlier.
0 The gas-exchange abnormalities in the setting of worsen-
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 ing airflow obstruction are several. First, although there
FIO2 may be a transient hyperventilation due to dyspnea in the
asthmatic patient, the worsening respiratory failure in
Targets: PO2 55-80, Pplat  30-35
obstructive lung disease is generally characterized by a
Figure 101-5  Two positive end-expiratory pressure (PEEP)/FIO2 algo- falling minute ventilation as inspiratory muscles fatigue in
rithms that have been used in National Institutes of Health Acute Res- the face of airflow obstruction. The result is termed hyper-
piratory Distress Syndrome Network clinical trials. Plotted are FIO2 on
the horizontal axis and PEEP on the vertical axis. With both approaches,
capnic respiratory failure. Second, as noted earlier, regional
the oxygenation target (PaO2 55 to 80 mm Hg, SPO2 88% to 95%) and the lung compression and regional hypoventilation produce
maximal allowable plateau pressure (35 cm H2O) is the same. Patients   mismatch that results in progressive hypoxemia. Alve-
V/Q
are moved in steps up and down the algorithm according to these targets. olar inflammation and flooding, however, are not charac-
The two algorithms depicted either focus on higher PEEP levels (“high teristic features of respiratory failure due to pure airflow
PEEP”—brown bars) or lower PEEP levels (“low PEEP”—blue bars).
(Courtesy Art Wheeler, MD, personal communication and Ref. 70.) obstruction, and thus shunts are less of an issue than in
parenchymal lung injury. Third, overdistended regions of
the lung coupled with underlying emphysematous changes
O2 saturation. Although most of the reported PEEP/FIO2 in some patients result in capillary loss and increasing dead
algorithms generally target modest levels of oxygenation space. This wasted ventilation further compromises the
(e.g., arterial PO2 55 to 80 mm Hg or SO2 88% to 95%), they ability of the inspiratory muscles to supply an adequate
tend to segregate into approaches that focus either on ventilation for alveolar gas exchange. These emphysema-
higher PEEP levels (“high PEEP” algorithms) or lower PEEP tous regions also have low recoil properties, which can
levels (“low PEEP” algorithms) (Fig. 101-5). Several trials worsen air trapping. Fourth, hypoxemic pulmonary vaso-
have compared high versus low PEEP algorithms in con- constriction coupled with chronic pulmonary vascular
junction with low tidal volume/limited Pplat strategies in changes in some airway diseases overload the right ventri-
patients with ARDS.119-121 A meta-analysis of trials with cle, further decreasing blood flow to the lung and increasing
mild ARDS (arterial PO2/FIO2 > 200) compared with more dead space further.
severe ARDS (arterial PO2/FIO2 < 200) demonstrated that Frequency-tidal volume settings in obstructive diseases
higher PEEP strategies had a significant mortality benefit in are chosen on the basis of many considerations similar to
more severe ARDS, whereas low PEEP strategies showed a those in parenchymal lung injury. Specifically, tidal volumes
trend toward a potential benefit in mild ARDS.122 should be in the 6 to 8 mL/kg (IBW) range with Pplat
In patients with abnormal chest wall mechanics (i.e., stiff targets of less than 30 cm H2O.67-70 In obstructive diseases,
chest walls from obesity, anasarca, abdominal compart- however, clinicians should also be aware that high peak
ment syndrome, and even systemic inflammation), the reli- airway pressures, even in the presence of acceptable values
ance on circuit airway pressures to guide ventilator settings for Pplat, may transiently subject regions of the lung to
will ignore the effect of elevated pleural pressures on reduc- periods of overdistention injury (see Fig. 101-2). Finally,
ing transpulmonary pressure, the ultimate determinant of setting the VT should also take into consideration the poten-
VILI and alveolar recruitment. Under these circumstances, tial to develop PEEPi with its consequent impact on Pplat
clinicians should consider modest empirical increases in elevations and the risk of overdistention injury.
applied PEEP and allowances of an increased Pplat beyond The ventilatory rate is used to control pH. Unlike the situ-
the limits described earlier. Alternatively, an esophageal ation in parenchymal disease, however, the elevated airway
catheter to measure Pes could be inserted to assess trans- resistance (and often the low recoil pressures of emphy-
pulmonary pressure directly and adjust settings accord- sema) greatly increase the potential for PEEPi and thus limit
ingly.19,123 Indeed, one clinical trial using Pes to guide PEEP the range of breath rates available. Indeed, reductions in
settings suggested improved outcomes.123 both VT and ventilatory rate to levels resulting in hypoven-
tilation and “permissive (or even “therapeutic”78 ) hyper-
Obstructive Airway Disease capnia” may be an appropriate tradeoff to limit PEEPi
Increases in airway resistance lead to respiratory failure development and overdistention.
from airflow obstruction and create two important patho- The I : E ratio in obstructive lung disease is generally set
physiologic changes. First, the increased pressures required as low as possible to minimize the development of air trap-
for airflow may overload inspiratory muscles, producing a ping. For the same reason, approaches using IRV strategies
“ventilatory pump failure” in which spontaneous minute are almost always contraindicated.
ventilation becomes inadequate for gas exchange. Second, The PEEP/FIO2 settings assume a different role in obstruc-
the narrowed airways create regions of lung that cannot tive lung disease than in parenchymal lung disease. Because
properly empty and return to their normal “resting volume,” alveolar recruitment is less of an issue and overdistention
thereby producing PEEPi.27 These regions of overinflation is more of an issue in obstructive lung injury than in paren-
create dead space and put inspiratory muscles at a substan- chymal lung injury, the strategy in using the PEEP/FIO2
101  •  Mechanical Ventilation 1773

steps in Figure 101-5 should probably be shifted more risk of premature withdrawal with consequent loss of
toward the use of FIO2 for oxygenation support instead of airway patency, aspiration, and inspiratory muscle fatigue.
PEEP. A specific role for PEEP in the obstructed patient arises An evidence-based task force125 has recommended a two-
when PEEPi serves as an added inspiratory threshold load step process:
on the patient attempting to trigger a breath, as described 1. Consider a patient a candidate for withdrawal if (a) the
earlier. Under these conditions, judicious application of lung injury is stable/resolving, (b) the gas exchange is
circuit PEEP (at levels up to 75% to 85% of PEEPi) can adequate with low PEEP/FIO2 requirements, (c) the
“counterbalance” PEEPi throughout the ventilator cir- hemodynamics are stable without pressors, and (d) the
cuitry to reduce this triggering load and facilitate the trig- patient has the capability to initiate spontaneous breaths.
gering process.87,88 2. In these patients, perform a spontaneous breathing trial
In severe airflow obstruction, the use of the low-density (using T-piece, CPAP, or pressure support at 5 cm H2O)
gas helium can help facilitate ventilation. Helium mixtures for 30 to 120 minutes. Assessments should include the
are is available as 80 : 20, 70 : 30, or 60 : 40 helium : oxygen ventilatory pattern, gas exchange, hemodynamics, and
breathing gas mixtures (heliox) and can both reduce patient comfort. Patients “passing” this trial should be consid-
inspiratory work and facilitate lung emptying (recall that ered for ventilator withdrawal.
driving pressure decreases and/or flow increases as gas
density decreases).124 However, to date there have been no In patients passing the spontaneous breathing trial (SBT),
studies demonstrating improved outcomes with the use of separate assessments are required to determine if the arti-
heliox. With a helium : oxygen gas mixture, it should be ficial airway can be removed. These involve cough strength,
remembered that many flow sensors must be recalibrated suctioning frequency, and, to a certain extent, the ability to
to account for the change in gas density. follow commands.125 Cuff leak, a bedside test that may dem-
Finally, it is important to note that noninvasive ventila- onstrate whether there is airway edema or compression
tion has been demonstrated most convincingly to improve around the endotracheal tube that could lead to postextu-
outcomes in patients with obstructive diseases.125 In setting bation airway obstruction, does not appear to be an impor-
up noninvasive ventilation, the same principles described tant predictor of success, except perhaps in the setting of
previously for invasive mechanical ventilation also apply prior upper airway injury. Extubation failures can be
(see Chapter 102). expected in 10% to 20% of all extubations. Many of these
involve airway protection issues and thus indicate the
Neuromuscular Respiratory Failure need for prompt reintubation. However, in some patients,
The risk of VILI is generally less in a patient with neuromus- especially those with chronic obstructive pulmonary
cular failure (e.g., central nervous system injury, drug disease, an extubation failure caused by increasing inspira-
overdoses, anesthesia) because lung mechanics are often tory muscle overload might be managed by noninvasive
near normal and regional overdistention less likely. More ventilation.127,128
“generous” VTs (e.g., up to 10 mL/kg PBW) have thus been In patients failing the SBT, a stable and comfortable level
proposed as being useful to improve comfort, maintain of support should be provided until the next SBT.125 Fre-
recruitment, prevent atelectasis, and avoid hypercarbia that quent (e.g., every 2 to 12 hours) support reductions are
may adversely affect central nervous system function. usually not necessary because they have not been shown to
However, this notion has been challenged recently by clini- speed up the withdrawal process and only serve to consume
cal trials of patients with normal lungs in the perioperative resources and expose the patient to the risks of muscle over-
period showing reduced postoperative complications when load. Repeat assessments for SBTs, however, should be done
VTS are used in the 6 to 8 mL/kg (IBW) range. Regardless daily.125,126 Importantly, aggressive strategies to reduce
of VT settings, maximal distending pressures should be kept sedation can accelerate this process.129 Indeed, some have
as low as possible while still being compatible with the other advocated “spontaneous awakening trials” coupled to SBTs,
goals noted previously.67-70 but it is not clear if this approach is superior to carefully
Mode selection in neuromuscular disease patients is often targeted sedation protocols.130
determined by patient comfort and reliability of the respira- A common problem seen in patients recovering from
tory drive. PEEP, even at low levels, is often beneficial at respiratory failure but still deemed in need of an artificial
preventing derecruitment (atelectasis) in these patients, airway is the presence of large VTs despite low levels of
who are often supine and unable to clear secretions or sigh. ventilatory support (e.g., inspiratory pressure levels of 5 cm
H2O). Under these circumstances, a search should be made
Recovering Respiratory Failure—“Weaning” and for causes of an excessive respiratory drive such as pain,
Discontinuation Process metabolic acidosis, or anxiety, and these issues should be
As respiratory failure stabilizes and begins to reverse, clini- addressed. Most would argue that, in the absence of revers-
cal attention shifts to the process of ventilator withdrawal. ible causes, sedation should not be used simply to lower
Unfortunately, a number of large clinical trials have clearly the VT.
demonstrated that current assessment/management strat- Regardless of the clinical scenario, whenever assisted-
egies are not optimal, resulting in considerable delay in ven- supported breaths are used, attention must be paid to assur-
tilator withdrawal.125,126 Such delay leads to increased ing synchronous interactions. As noted earlier, this first
length of ICU stay, increased costs, prolonged exposure to means addressing the appropriateness of respiratory drive
circuit pressure, and increased risk of infection. Attempts to ensure that reversible causes of excess drive (e.g., pain,
to speed withdrawal, however, must be balanced against the anxiety, acidosis) are managed and then providing settings
1774 PART 3  •  Clinical Respiratory Medicine

that maximize trigger sensitivity. It also ensures proper flow HFV can be supplied by either jets or oscillators. Jets
and cycle synchrony. inject high-frequency pulses of gas into the airways. Oscil-
lators literally vibrate a fresh bias flow of gas delivered at
the tip of the endotracheal tube. Because of this, oscillatory
HFV has sometimes been referred to as “CPAP with a
RECENT INNOVATIONS wiggle.”
IN MECHANICAL The putative advantages of HFV are twofold. First, the
VENTILATORY SUPPORT small alveolar tidal pressure swings minimize cyclical over-
distention and derecruitment. Second, a high mean airway
INNOVATIVE STRATEGIES FOR pressure can also prevent derecruitment. Interestingly,
mean pressures used during HFV are often reported to
“LUNG PROTECTION”
exceed the 30 to 35 cm H2O threshold employed during
Several innovations introduced recently may assist clini- conventional ventilation. This tolerance of a higher mean
cians in reducing ventilator-induced lung injury. Among pressure with HFV may be explained by a better maintained
the more interesting are airway pressure release ventilation alveolar structure with a slowly applied (albeit vibrating)
(APRV) and high-frequency ventilation (HFV). In addition, constant pressure as opposed to cyclical brief tidal
ventilating patients in the prone position, although not pressures.133
novel, can also be viewed as an interesting strategy to facili- Clinical experience with HFV has been most extensive in
tate lung protective ventilation. the neonatal and pediatric age groups where a number of
trials show improved long-term lung function when HFV is
Airway Pressure Release Ventilation used in acute respiratory failure.134,135 Although adult expe-
APRV (also known as biphasic ventilation, bilevel ventilation, rience with HFV is less, a meta-analysis of randomized trials
and bilevel positive airway pressure) is a time-cycled, pressure- using HFV in adult respiratory failure suggested an outcome
targeted form of ventilatory support.111-115 It is actually a benefit to HFV.136 Two large subsequent trials, however,
modification of pressure-targeted SIMV that allows sponta- have called this conclusion into question. In one, mortality
neous breathing (with or without PS) during both the infla- was identical in both the HFV and lung-protective conven-
tion and deflation phases. tional strategy137; in the other, the trial was stopped early
The putative advantages of this approach are similar to because of increased mortality in the HFV group.138 Of
others that utilize long inspiratory times. Specifically, the concern in both of these trials was the use of many centers
long inflation phase recruits more slowly filling alveoli and that had little or no previous experience with HFV. Never-
raises mean airway pressure without increasing applied theless, both trials make it clear that HFV, if used, is best
PEEP (although PEEPi can develop with short deflation reserved for patients failing conventional lung protective
periods). Unlike older IRV strategies that required paralysis, strategies and should be employed by clinicians experienced
however, the additional spontaneous efforts during infla- in the use of HFV.
tion may enhance both recruitment and cardiac filling
when compared with other controlled forms of support. Positive-Pressure Ventilation in
Although IRV strategies are usually reserved for severe the Prone Position
forms of respiratory failure in which airway pressures and Mechanical ventilation of patients in the prone position
FIO2 levels are approaching potentially injurious levels, the offers several physiologic advantages.139 Chief among
comfort and recruitment potential associated with APRV these are two mechanisms that can improve ventilation
may prompt consideration of its use even in less severe distribution. First, the heart no longer rests on the left
forms of lung injury. lower lobe and this allows better ventilation distribution
Good gas exchange has been demonstrated with APRV in to that region. Second, the sternum is restricted in its
several small observational clinical trials, often with lower ability to move outward. This functionally stiffens the
maximal airway pressures than used with control ventila- chest wall and forces a more even distribution of positive
tion.111 However, the end-inspiratory lung distention in pressure.
APRV may not be less than that provided during other Prone positioning has its challenges, primarily from a
forms of support (and, indeed, it could be substantially nursing perspective. The act of proning requires careful
higher) because spontaneous VTs can expand the lung teamwork (although automated proning beds exist). More
beyond the volume at the APRV set pressure. In the few importantly, management of vascular lines, feeding tubes,
randomized studies that compare APRV to a true lung- artificial airways, ostomies, and so on need careful atten-
protective strategy, no differences in important outcomes tion because they can be easily dislodged. Avoiding facial
have been found.113-115 pressure sores also requires careful nursing care.
There have been a number of randomized controlled
High-Frequency Ventilation trials of proning in patients with ARDS and, until recently,
HFV uses high breathing frequencies (120 to 900 breaths/ despite consistently better oxygenation, no consistent
min in the adult) coupled with small VTs (usually < ana- outcome benefit could be demonstrated.139 However, the
tomic dead space and often < 1 mL/kg IBW at the alveolar largest study to date focusing on ARDS patients with arte-
level) to provide gas exchange in the lungs.131 Gas transport rial PO2/FIO2 ratios less than 150 recently showed a signifi-
under these seemingly unphysiologic conditions may cant mortality benefit to lung protective mechanical
involve such mechanisms as Taylor dispersion, coaxial ventilation provided in the prone position for more than 16
flows, and augmented diffusion.132 hours/day.140
101  •  Mechanical Ventilation 1775

AUTOMATED WEANING STRATEGIES scenarios, however, may be amenable to these automated


approaches: The first is the patient rapidly recovering from
Over the years, a number of attempts have been made to sedatives/anesthesia where these systems can alert clini-
“automate” the weaning process.7 An early example is cians to the return of adequate spontaneous efforts. The
minimum minute ventilation, which adjusted the intermit- second is the patient requiring prolonged mechanical ven-
tent mandatory breath rate according to the level of spon- tilation who has failed multiple SBTs. Under these circum-
taneous ventilation. The concept behind automated stances, an automated support reduction system could
weaning was that significant clinician time could be saved serve as a diagnostic tool alerting the clinician to recovery
and ventilatory support could be automatically reduced of respiratory function and the possibility of reinstating
in a timely fashion on the basis of simple ventilator SBTs.
measurements.
Volume support (VS, also known by trade names such as OPTIMIZING SYNCHRONY DURING
“Automatic Pressure Ventilation”) is a newer strategy with
INTERACTIVE BREATHS
the potential for automatic support reduction.7 As described
earlier, VS is a pressure support mode that uses tidal volume Interactive breaths are commonly used during mechanical
as a feedback control for continuously adjusting the pres- ventilatory support to improve comfort (and reduce seda-
sure support level. Proponents claim that this approach tion), especially during the recovery phase of respiratory
could “automatically” wean a patient by reducing pressure failure. As noted previously, interactive breaths need to be
support as patient effort increases and respiratory system synchronous with patient efforts during all three phases of
mechanics improve. Conversely, pressure support would breath delivery: trigger, flow delivery, and cycle. A number
increase if patient effort diminished or respiratory system of recent innovations have been introduced and are
mechanics worsened. Similarly, it has also been suggested reviewed here.
that VS may be a useful way to maintain a more constant Although all of these innovations have conceptual appeal
level of partial support in patients with fluctuating levels of and have been shown to perform as designed in both bench
effort related to drugs or neurologic conditions. All of these testing and small clinical observational trials, patient out-
effects have been demonstrated in small studies focused on comes, including sedation needs, ventilator days, or patient
patients with rapidly recovering respiratory failure.7 comfort assessments, have generally not been studied. Nev-
Unfortunately, the simplicity of VS may introduce prob- ertheless, their straightforward designs, ease of operation,
lems.7,26 For instance, if the volume set by the clinician is and safety make these innovations appropriate to consider
excessive for patient demand, a recovering patient may not in patients receiving interactive breaths.
attempt to take over the work of breathing for that volume
and thus weaning may not progress. In addition, if the Endotracheal Tube Resistance Compensation
pressure level increases in an attempt to maintain an The endotracheal tube provides a significant resistance to
inappropriately high-set VT in the patient with airflow flow during both inspiration and expiration. During the
obstruction, PEEPi may result. Conversely, if the volume inspiratory phase, this means that pressure build-up in the
set by the clinician is not adequate for patient demand, airways “lags” behind the pressure build-up in the ventila-
a patient may not receive adequate support. Under these tor circuitry. Thus the “square” wave of pressure in the
conditions, a patient will perform excessive work to main- circuitry provided by a pressure-targeted breath is distorted
tain a certain VT even as the inspiratory pressure is being in the airways to a slower rise of pressure. This may create
reduced. A transient increase in patient demand from pain significant initial flow dys-synchrony in patients with vigor-
or anxiety could also result in inappropriate support reduc- ous inspiratory efforts. During expiration, a similar gradi-
tion with VS.26 ent between airway pressures and set circuit PEEP can
Adaptive support ventilation (ASV) has a sophisticated develop.
algorithm to adjust the ventilatory pattern during machine- One way to address this is to target ventilator pressures
triggered breaths as described earlier. However, ASV per- to a measured tracheal pressure distal to the endotracheal
forms similarly to VS during patient-triggered breaths. tube. Unfortunately, readings from intra-airway pressure
Small clinical studies have demonstrated that ASV can sensors over prolonged periods are not reliable. Another
automatically wean ventilatory support safely.141 However, approach is to account for endotracheal tube resistance
large trials comparing ASV weaning to regular daily SBT mathematically in the ventilator flow delivery pattern.145,146
strategies have not been done. Known by various trade names (e.g., “Automatic Airway
Another commercially available feedback system for Compensation,” “Automatic Tube Compensation”), this
weaning pressure-targeted breaths uses not only VT but approach initially provides an inspiratory pressure higher
also respiratory rate and end-tidal carbon dioxide to adjust than the set pressure target. As inspiration proceeds, this
ventilator settings.142 Randomized trials with this system, delivered pressure then tapers to the set inspiratory pres-
however, have not shown faster ventilator withdrawal when sure target. This compensation mechanism can also operate
compared with strategies using protocols for regular in expiration with an initial expiratory airway pressure
SBTs.143 below the set PEEP that then rises to the set PEEP. A more
Inherent in all of these automated weaning systems is the square wave pattern of inspiratory and expiratory tracheal
notion that gradual support reductions (weaning) in pressures is the result.145,146
between SBTs facilitates ventilator withdrawal—a notion Applying endotracheal tube resistance compensation
that has little, if any, evidence supporting it in patients is relatively straightforward. Clinicians must input the
recovering from acute respiratory failure.144 Two clinical characteristics of the endotracheal tube. Thereafter, the
1776 PART 3  •  Clinical Respiratory Medicine

ventilator provides the appropriate circuit pressure profile ized by a comfortable patient without evidence on the
during both inspiration and expiration to create the desired circuit pressure graph of continued inspiratory effort after
square wave pattern in the trachea. Although outcome cycling (premature cycling) or of expiratory efforts begin-
studies using this compensation approach have not been ning during the inspiratory phase (delayed cycling).
done, the conceptual appeal should make it a consideration Although no outcome studies have been performed using
in virtually all patients receiving assisted/supported these cycle adjusters, their physiologic appeal, ease of use,
pressure-targeted breaths—especially those with vigorous and apparent safety should make them a consideration in
inspiratory efforts. virtually every patient receiving pressure support.
Pressure Rate of Rise (Slope) Adjusters Proportional Assist Ventilation
The original design for pressure-targeted breaths (pressure Proportional assist ventilation (PAV) is a novel approach to
support and pressure assist-control) had a programmed assisted ventilation in which the clinician sets the “gain” on
flow delivery algorithm that attempted to reach the target patient-generated flow and volume.149-150 In PAV, there is
inspiratory pressure quickly, without causing an uncom- thus no set pressure, flow, or volume. Instead, the sensed
fortable overshoot. Newer ventilators, however, allow the patient effort is boosted according to a proportion of the
clinicians to adjust the rate of rise of this pressure (slope measured work of breathing set by the clinician.
adjusters), and clinical studies have suggested that slope PAV requires that “test breaths” (controlled breaths with
adjustment could significantly enhance flow synchrony in fixed flow and volume) be given. This allows for the calcula-
many patients.147 Specifically, these studies found that a tion of respiratory system mechanics, which can be coupled
rapid rate of rise was often desired in a patient with vigorous with the measured ventilation to calculate work of breath-
flow demands, whereas a much slower rate of rise was often ing (resistive and elastic ventilatory muscle loads). These
preferable in patients with less vigorous demands. load calculations are repeated at regular intervals in order
There are several approaches to setting the slope adjuster. to maintain reliable inputs for the PAV algorithm.
The most direct way is to use the circuit pressure graph and PAV breaths are patient-initiated breaths triggered in a
adjust the slope to create a “smooth square wave” appear- conventional way using circuit pressure or flow sensors.
ance to the circuit pressure profile. Studies have also shown Thereafter, the ventilator continues to monitor flow and
that an optimal slope setting correlates with the greatest VT volume demanded by the patient and puts a clinician-set
for a given pressure setting.147 Patient comfort should “gain” on this demand to augment flow and pressure in
always be considered in determining optimal slope proportion to the desired reduction in the patient’s work of
settings. breathing. The PAV breath cycles when sensed flow demand
has ceased.
Pressure Support Cycle Adjusters Like pressure-targeted breaths, PAV flow delivery varies
Pressure support breaths have a flow cycling mechanism to with patient effort; unlike pressure-targeted breaths, pres-
terminate the breath. On earlier machines, a set flow termi- sure also varies with patient effort. The conceptual upside
nation criterion was usually set by the manufacturer (e.g., to PAV is that flow and cycle synchrony should be enhanced
25% to 35% of peak flow). Although this set flow was often over conventional flow- or pressure-targeted breaths.
effective, it could sometimes terminate breaths too early in Another conceptual upside is that patient-driven tidal
patients with long inspiratory demands and it could some- volume variability with its theoretical lung protective ben-
times terminate too late, typically in patients with obstruc- efits may be enhanced.151 The downside, however, is that,
tive airway disease. In this latter situation, air trapping unlike conventional pressure-targeted breaths, there is no
could also be made worse because of the resulting shorter minimum pressure or flow provided. Thus PAV must be used
expiratory time. with caution in patients with unreliable ventilatory drives
There are several approaches to improving cycle syn- from either disease or drugs. Indeed, with all patients on
chrony with pressure support. One is to switch from pres- PAV, careful monitoring and backup support modes should
sure support to a pressure-assist breath (patient-triggered, be available.
pressure-targeted, time-cycled breath as is usually available Most clinical studies with PAV have shown enhanced syn-
on most machines providing pressure ACV if the set rate is chrony compared with conventional modes.149-156 However,
turned low or off). This breath provides direct clinician it is not clear what the ideal PAV gain(s) should be in various
control of inspiratory time and thus of cycling. Another clinical settings. Moreover, to date, there have been no good
strategy is to adjust the pressure slope setting described pre- randomized trials looking at important outcome benefits
viously on pressure support breaths. A rapid peak initial (e.g., ventilator duration, sedation needs, mortality) when
flow will have a correspondingly high flow cycle variable PAV is compared with conventional assisted/supported
(and thus short inspiratory time); a very slow peak initial ventilation.
flow will have a correspondingly low flow cycle variable
(and thus long inspiratory time). Neurally Adjusted Ventilatory Assistance
A newer approach is to allow adjustments of the flow Neurally adjusted ventilatory assistance (NAVA) utilizes the
criteria of the pressure support cycle to assure appropriate diaphragmatic electromyographic (EMG) signal to trigger,
synchrony of the cycle with the end of patient effort.148 As adjust flow, and cycle assisted breaths.153,157 NAVA requires
with other adjustments of interactive breaths, airway pres- placement of a unique esophageal catheter with an array
sure graphics and assessments of patient comfort should of diaphragm EMG sensors. These sensors detect the onset,
guide adjustments. Proper breath synchrony is character- intensity, and termination of inspiratory efforts directly.
101  •  Mechanical Ventilation 1777

Like PAV, a clinician-set gain is then applied, which deter- Complete reference list available at ExpertConsult.
mines flow and pressure delivery in proportion to the EMG
signal. Key Readings
The conceptual benefit to NAVA is that synchrony with ACCP/AARC/SCCM Task Force: Evidence based guidelines for weaning
all three phases of breath delivery (trigger, gas delivery, and and discontinuing mechanical ventilatory support. Chest 120(Suppl 6),
2001.
cycle) should be enhanced over conventional flow- or Brander L, Ranieri VM, Slutsky AS: Esophageal and transpulmonary pres-
pressure-targeted breaths. Like PAV, another conceptual sure help optimize mechanical ventilation in patients with acute lung
benefit is that patient-driven tidal volume variability with injury. Crit Care Med 34(5):1556–1558, 2006.
its theoretical lung protective benefits may be enhanced. Briel M, Meade M, Mercat A, et al: Higher vs lower positive end-expiratory
Also like PAV, the downside is that there is no minimum pressure in patients with acute lung injury and acute respiratory dis-
tress syndrome: systematic review and meta-analysis. JAMA 303(9):
pressure or flow provided. Thus, like PAV, NAVA must be 865–873, 2010.
used with caution in patients with unreliable ventilatory Brower RG, Hubmayr RD, Slutsky AS: Lung stress and strain in acute
drives from either disease or drugs. Moreover, with NAVA, respiratory distress syndrome: good ideas for clinical management? Am
there is also concern about the stability of the EMG signal J Respir Crit Care Med 178(4):323–324, 2008.
Chatburn RL, Branson RD: Classification of mechanical ventilators.
coming from a catheter that can move within the esopha- In MacIntyre NR, Branson RD, editors: Mechanical ventilation, ed 2,
gus. Thus all patients on NAVA require careful monitoring St. Louis, 2009, Saunders Elsevier, pp 1–49.
and backup support modes. Cooke CR, Kahn JM, Watkins TR, et al: Cost effectiveness of implementing
Most clinical studies with NAVA have shown enhanced low tidal volume ventilation in patients with acute lung injury. Chest
synchrony compared with conventional modes.158-161 136:79–88, 2009.
Dodek P, Keenan S, Cook D, et al: Evidence-based clinical practice guideline
However, like PAV, it is unclear what the optimal EMG gain for the prevention of ventilator-associated pneumonia. Ann Intern Med
setting(s) should be in various clinical settings. To date 141:305–313, 2004.
there have been no good randomized trials looking at impor- Ferrer M: Non-invasive ventilation in the weaning process. Minerva Anes-
tant outcome benefits (e.g., ventilator duration, sedation tesiol 74:311–314, 2008.
Gattinoni L, Caironi P, Cressoni M, et al: Lung recruitment in patients with
needs, mortality) when NAVA is compared with conven- the acute respiratory distress syndrome. N Engl J Med 354:1775–1786,
tional assisted/supported ventilation. 2006.
Gattinoni L, Pelosi P, Crotti S, et al: Effects of positive end expiratory pres-
sure on regional distribution of tidal volume and recruitment in adult
respiratory distress syndrome. Am J Respir Crit Care Med 151:1807–
Key Points 1814, 1995.
Gilstrap D, MacIntyre NR: Patient-ventilator interactions. Am J Resp Crit
■ Positive-pressure breaths are characterized by three Care Med 188:1058–1068, 2013.
variables: the breath trigger, the flow delivery target Guérin C, Reignier J, Richard JC, et al: PROSEVA Study Group. Prone posi-
(pressure or flow), and the cycle criteria. tioning in severe acute respiratory distress syndrome. N Engl J Med
■ The interactions of positive-pressure breaths and 368(23):2159–2168, 2013.
Lellouche F, Brochard L: Advanced closed loops during mechanical ventila-
respiratory system mechanics are described by the tion (PAV, NAVA, ASV, SmartCare). Clin Anaesthesiol 23:81–93, 2009.
equation of motion: Pressure = (flow × resistance) + Mehta S, Burry L, Cook D, et al: SLEAP Investigators; Canadian Critical
(volume/system compliance). Care Trials Group. Daily sedation interruption in mechanically venti-
■ In lungs with collapsed alveoli, alveolar recruitment is lated critically ill patients cared for with a sedation protocol: a random-
ized controlled trial. JAMA 308(19):1985–1992, 2012.
accomplished by transient increases in positive- Slutsky AS, Ranieri VM: Ventilator-induced lung injury. N Engl J Med
pressure inflation and is maintained by positive end- 369(22):2126–2136, 2013.
expiratory pressure. Tremblay LN, Slutsky AS: Ventilator-induced lung injury: from the bench
■ Ventilator-induced lung injury can be caused by to the bedside. Intensive Care Med 32(1):24–33, 2006.
several mechanisms, including alveolar overdistention Truwit JD, Marini JJ: Evaluation of thoracic mechanics in the ventilated
patient. Part I: primary measurements. J Crit Care 3:133–150, 1988.
and repetitive alveolar collapse/reopening. Truwit JD, Marini JJ: Evaluation of thoracic mechanics in the ventilated
■ Removing mechanical ventilatory support expedi- patient. Part II: applied mechanics. J Crit Care 3:192–213, 1988.
tiously requires regular sedation adjustments and Yilmaz M, Keegan MT, Iscimen R, et al: Toward the prevention of acute
assessments of spontaneous breathing capabilities. lung injury: protocol-guided limitation of large tidal volume ventilation
and inappropriate transfusion. Crit Care Med 35:1660–1666, 2007.
101  •  Mechanical Ventilation 1777.e1

References 27. Marini JJ, Crooke PS: A general mathematical model for respiratory
dynamics relevant to the clinical setting. Am Rev Respir Dis 147:14–
1. MacIntyre NR: Mechanical ventilation: the next 50 years. Respir 24, 1993.
Care 43:490–493, 1998. 28. Macklen PT: Relationship between lung mechanics and ventilation
2. ANZIC Influenza Investigators: Critical care services and 2009 distribution. Physiology. 16:580–588, 1973.
H1N1 influenza in Australia and New Zealand. N Engl J Med 29. Mili-Emili J, Henderson JAN, Dolovich MB, et al: Regional distribu-
361:1925–1934, 2009. tion of inhaled gas in the lung. J Appl Physiol 21:749–759, 1966.
3. Mushin M, Rendell-Baker W, Thompson PW, Mapleson WW: 30. Abraham E, Yoshihara G: Cardiorespiratory effects of pressure con-
Automatic ventilation of the lungs, Oxford, UK, 1980, Blackwell, trolled ventilation in severe respiratory failure. Chest 98:1445–
pp 62–160. 1449, 1990.
4. American Society for Testing and Materials: Standards specifications 31. Pratt PC: Pathology of the adult respiratory distress syndrome. In
for ventilators intended for use in critical care (ASTM Standards No. 36), Thurlbeck WM, Ael MR, editors: The lung: structure, function and
Philadelphia, 1991, American Society for Testing and Materials, disease, Baltimore, 1978, Williams & Wilkins, pp 43–57.
pp 1123–1155. 32. AARC Conference on Positive End Expiratory Pressure. Respir Care
5. Chatburn RL, Branson RD: Classification of mechanical ventilators. 33:419–527, 1988.
In MacIntyre NR, Branson RD, editors: Mechanical ventilation, ed 2, 33. Gattinoni L, Pesenti A, Baglioni S, et al: Inflammatory pulmonary
St. Louis, 2009, Saunders Elsevier, pp 1–49. edema and PEEP: correlation between imaging and physiologic
6. American Association for Respiratory Care Consensus Group: Essen- studies. J Thorac Imaging 3:59–64, 1988.
tials of mechanical ventilators. Respir Care 37:1000–1008, 1992. 34. Gattinoni L, Pelosi P, Crotti S, et al: Effects of positive end expiratory
7. Branson RD, MacIntyre NR: Dual control modes of mechanical ven- pressure on regional distribution of tidal volume and recruitment in
tilation. Respir Care 41:294–305, 1996. adult respiratory distress syndrome. Am J Respir Crit Care Med
8. Campbell RS, Branson RD, Johannigman JA: Adaptive support ven- 151:1807–1814, 1995.
tilation. Respir Care Clin N Am 7:425–440, 2001. 35. Gattinoni L, Caironi P, Cressoni M, et al: Lung recruitment in patients
9. Kondili E, Xirouchaki N, Georgopoulos D: Modulation and treatment with the acute respiratory distress syndrome. N Engl J Med
of patient-ventilator dys-synchrony. Curr Opin Crit Care 13:84–89, 354:1775–1786, 2006.
2007. 36. Lim SC, Adams AB, Simonson DA, et al: Intercomparison of recruit-
10. Tom LR, Sasoon CSH: Patient ventilator interactions. In MacIntyre ment maneuver efficacy in three models of acute lung injury. Crit
NR, Branson RD, editors: Mechanical ventilation, ed 2, St. Louis, Care Med 32:2371–2377, 2004.
2009, Saunders Elsevier, pp 182–197. 37. Armstrong BW, MacIntyre NR: Pressure controlled inverse ratio ven-
11. Gilstrap D, MacIntyre NR: Patient-ventilator interactions. Am J Resp tilation that avoids air trapping in ARDS. Crit Care Med 23:279–285,
Crit Care Med 188(9):1058–1068, 2013. 1995.
12. Sassoon CSH: Mechanical ventilator design and function: the trigger 38. Cole AGH, Weller SF, Sykes MD: Inverse ratio ventilation compared
variable. Respir Care 37:1056–1069, 1992. with PEEP in adult respiratory failure. Intensive Care Med 10:227–
13. Branson RD: Humidification and aerosol therapy. In MacIntyre NR, 232, 1984.
Branson RD, editors: Mechanical ventilation, ed 2, St. Louis, 2009, 39. Tharratt RS, Allen RP, Albertson TE: Pressure controlled inverse
Saunders Elsevier, pp 111–124. ratio ventilation in severe adult respiratory failure. Chest 94:755–
14. Pinsky MR, Hrehocik D, Culpepper JA, Snyder JV: Flow resistance of 762, 1998.
expiratory positive pressure systems. Chest 94:788–791, 1998. 40. Webb HJH, Tierney DF: Experimental pulmonary edema due to inter-
15. Pepe PE, Marini JJ: Occult positive end-expiratory pressure in mittent positive pressure ventilation with high inflation pressures:
mechanically ventilated patients with airflow obstruction. Am Rev protection by positive end-expiratory pressure. Am Rev Respir Dis
Respir Dis 126:166–170, 1982. 110:556–565, 1974.
16. Hill NS, Brennan J, Garpestad E, Nava S: Noninvasive ventilation in 41. Manzano F, Fernandez-Mondejar E, Colmenero M, et al: Positive end
acute respiratory failure. Crit Care Med 35:2402–2407, 2007. expiratory pressure reduces incidence of ventilator-associated pneu-
17. Dhand R, Guntur VP: How best to deliver aerosol medications to monia in nonhypoxemic patients. Crit Care Med 36:2225–2231,
mechanically ventilated patients. Clin Chest Med 29:277–296, 2008.
2008. 42. Wyszogrodski I, Kyei-Aboagye K, Taaeusch HW Jr, Avery ME: Sur-
18. Marini JJ: What derived variables should be monitored during factant inactivation by hyperventilation: conservation by end-
mechanical ventilation? Respir Care 37:1097–1107, 1992. expiratory pressure. J Appl Physiol 38:461–466, 1975.
19. Brander L, Ranieri VM, Slutsky AS: Esophageal and transpulmonary 43. Grasso S, Stripoli T, De Michele M, et al: ARDSnet ventilatory proto-
pressure help optimize mechanical ventilation in patients with acute col and alveolar hyperinflation: role of positive end-expiratory pres-
lung injury. Crit Care Med 34(5):1556–1558, 2006. sure. Am J Respir Crit Care Med 176:761–767, 2007.
20. MacIntyre NR: Ventilator monitors and displays. In MacIntyre NR, 44. Ferragni PP, Rosbosh G, Tealdi A, et al: Tidal hyperinflation during
Branson RD, editors: Mechanical ventilation, ed 2, St. Louis, 2009, low tidal volume ventilation in acute respiratory distress syndrome.
Saunders Elsevier, pp 146–158. Am J Respir Crit Care Med 175:160–166, 2007.
21. Truwit JD, Marini JJ: Evaluation of thoracic mechanics in the venti- 45. MacIntyre NR, Leatherman NE: Mechanical loads on the ventilatory
lated patient. Part I: primary measurements. J Crit Care 3:133–150, muscles. Am Rev Respir Dis 144:968–973, 1989.
1988. 46. McGregor M: Bechlake MR. The relationship of oxygen cost of
22. Truwit JD, Marini JJ: Evaluation of thoracic mechanics in the venti- breathing to mechanical work and and respiratory force. J Clin Invest
lated patient. Part II: applied mechanics. J Crit Care 3:192–213, 40:971–980, 1961.
1988. 47. Bellemare F, Grassino A: Effect of pressure and timing of contrac-
22a.  Chiumello D, Cressoni M, Colombo A, et al: The assessment of tion on human diaphragm fatigue. J Appl Physiol 53:1190–1195,
transpulmonary pressure in mechanically ventilated ARDS patients. 1982.
Intensive Care Med 40(11):1670–1678, 2014. 47a.  Bellani G, Pesenti A: Assessing effort and work of breathing. Curr
23. Ranieri VM, Brienza N, Santostasi S, et al: Impairment of lung and Opini Crit Care 20(3):352–358, 2014.
chest wall mechanics in patients with acute respiratory distress syn- 48. Banner MJ, Kirby RR, MacIntyre NR: Patient and ventilator work of
drome: role of abdominal distension. Am J Respir Crit Care Med breathing and ventilatory muscle loads at different levels of pressure
156:1082–1091, 1997. support ventilation. Chest 100:531–533, 1991.
24. Chiumello D, Carlesso E, Cadringher P, et al: Lung stress and strain 49. Marini JJ: Exertion during ventilator support: how much and how
during mechanical ventilation for acute respiratory distress syn- important? Respir Care 31:385–387, 1986.
drome. Am J Respir Crit Care Med 178:346–355, 2008. 50. MacIntyre NR: Weaning from mechanical ventilatory support:
25. MacIntyre N, Marini JJ: Point-counterpoint: is pressure assist-control volume-assisting intermittent breaths versus pressure-assisting
preferred over volume assist-control mode for lung protective venti- every breath. Respir Care 33:121–125, 1988.
lation in patients with ARDS? Chest 140(2):286–294, 2011. 51. American Thoracic Society and European Respiratory Society:
26. Jaber S, Delay JM, Matecki S, et al: Volume-guaranteed pressure- Skeletal muscle dysfunction in chronic obstructive pulmonary
support ventilation facing acute changes in ventilatory demand. disease: a statement of the American Thoracic Society and European
Intensive Care Med 31:1181–1188, 2005. Respiratory Society. Am J Respir Crit Care Med 159:S1–S40, 1999.
1777.e2 PART 3  •  Clinical Respiratory Medicine

52. Vassilakopoulos D, Petrof B: Ventilator induced diaphragmatic dys- 75. Marini JJ, Hotchkiss JR, Broccard AF: Bench-to-bedside review:
function. Am J Respir Crit Care Med 169:336, 2004. microvascular and airspace linkage in ventilator-induced lung
53. Levine S, Nguyen T, Taylor N, et al: Rapid disuse atrophy of dia- injury. Crit Care (Lond) 7:435–444, 2003.
phragm fibers in mechanically ventilated humans. N Engl J Med 76. Vieira SR, Puybasset L, Lu Q, et al: A scinographic assessment of
358:1327–1335, 2008. pulmonary morphology in acute lung injury. Am J Resp Crit Care
54. Anzueto A, Peters JI, Tobin MJ, et al: Effects of prolonged controlled Med. 159:1612–1623, 1999.
mechanical ventilation on diaphragmatic function in healthy adult 77. MacIntyre NR: Supporting oxygenation in acute respiratory failure.
baboons. Crit Care Med 25:1187–1190, 1997. Respir Care 58(1):142–150, 2013.
55. Hansen-Flaschen J, Brazinsky S, Bassles C, Lanken PV: Use of sedat- 78. Ijland MM, Heunks LM, van der Hoeven JG, et al: Bench-to-bedside
ing drugs and neuromuscular blockade in patients requiring review: hypercapnic acidosis in lung injury—from “permissive” to
mechanical ventilation for respiratory failure. JAMA 266:2870– “therapeutic.” Crit Care 14(6):137–237, 2010.
2875, 1991. 79. Trembly L, Valenza F, Ribiero SP, et al: Injurious ventilatory strate-
56. Marini JJ, Culver BH, Butler J: Mechanical effect of lung inflation gies increase cytokines and C-fos mRNA expression in an isolated rat
with positive pressure on cardiac function. Am Rev Respir Dis lung model. J Clin Invest 99:944–952, 1997.
124:382–386, 1979. 80. Ranieri VM, Suter PM, Totorella C, et al: Effect of mechanical venti-
57. Scharf SM, Caldini P, Ingram RH Jr: Cardiovascular effects of lation on inflammatory mediators in patients with acute respiratory
increasing airway pressure in dogs. Am J Physiol 232:1135–1143, distress syndrome. JAMA 282:54–61, 1999.
1977. 81. Nahum A, Hoyt J, Schmitz L, et al: Effect of mechanical ventilation
58. Pinsky MR, Guimond JG: The effects of positive end-expiratory pres- strategy on dissemination of interacheally instilled E. coli in dogs.
sure on heart-lung interactions. J Crit Care 6:1–15, 1991. Crit Care Med 25:1733–1743, 1997.
59. Pinsky MR, Summer WR, Wise RA, et al: Augmentation of cardiac 82. Jenkinson SG: Oxygen toxicity. New Horizons. 1:504–511, 1993.
function by elevation of intrathoracic pressure. J Appl Physiol 83. Rachmale S, Li G, Wilson G, et al: Practice of excessive FIO2 and
54:450–455, 1983. effect on pulmonary outcomes in mechanically ventilated patients
60. Lemaire F, Teboul JL, Cinotti L, et al: Acute left ventricular dysfunc- with acute lung injury. Respir Care 57(11):1887–1893, 2012.
tion during unsuccessful weaning from mechanical ventilation. 84. Budinger GR, Mutlu GM: Balancing the risks and benefits of
Anesthesiology 69:171–179, 1988. oxygen therapy in critically ill adults. Chest 143(4):1151–1162,
61. Hughes JM, Glazier JB, Maloney JE, West JB: Effect of lung volume 2013.
on the distribution of pulmonary blood flow in man. Respir Physiol 85. Betbese AJ, Perez M, Bak E, et al: A prospective study of unplanned
4:58–72, 1968. endotracheal extubation in ICU patients. Crit Care Med 26:1180–
62. Anzueto A, Frutos-Vivar F, Esteban A, et al: Incidence, risk factors 1186, 1998.
and outcome of barotrauma in mechanically ventilated patients. 86. de Wit M, Miller KB, Green DA, Ostman HE: Ineffective triggering
Intensive Care Med 30:612–619, 2004. predicts increased duration of mechanical ventilation. Crit Care Med
62a.  Slutsky AS, Ranieri VM: Ventilator-induced lung injury. N Engl J 37:2740–2745, 2009.
Med 369:2126–2136, 2013. 87. Petrof BJ, Legare M, Goldberg P, et al: Continous positive airway pres-
62b.  Hasgemian SM, Mohajerani SA, Jamaati HR: Ventilator-induced sure reduces work of breathing and dyspnea during weaning from
lung injury. N Engl J Med (10):979–980, 2014. mechanical ventilation in severe chronic obstructive pulmonary
63. Kolobow T, Morentti MP, Fumagalli R, et al: Severe impairment in disease. Am Rev Respir Dis 141:281–289, 1990.
lung function induced by high peak airway pressure during mechan- 88. MacIntyre NR, McConnell R, Cheng KC: Applied PEEP reduces the
ical ventilation. Am Rev Respir Dis 135:312–315, 1987. inspiratory load of intrinsic PEEP during pressure support. Chest
64. Dreyfuss D, Savmon G: Ventilator induced lung injury: lessons 111:188–193, 1997.
from experimental studies. Am J Respir Crit Care Med 157:294–323, 89. Akoumianaki E, Lyazidi A, Rey N, et al: Mechanical ventilation-
1998. induced reverse-triggered breaths: a frequently unrecognized form
65. Tremblay LN, Slutsky AS: Ventilator-induced lung injury: from the of neuromechanical coupling. Chest 143(4):927–938, 2013.
bench to the bedside. Intensive Care Med 32(1):24–33, 2006. 90. Flick GR, Belamy PE, Simmons DH: Diaphragmatic contraction
66. Brower RG, Hubmayr RD, Slutsky AS: Lung stress and strain in acute during assisted mechanical ventilation. Chest 96:130–135, 1989.
respiratory distress syndrome: good ideas for clinical management? 91. Marini JJ, Smith TC, Lamb VI: External work output and force gen-
Am J Respir Crit Care Med 178(4):323–324, 2008. eration during synchronized intermittent mechanical ventilation.
67. Jia X, Malhotra A, Saeed M, et al: Risk factors for ARDS in patients Am Rev Respir Dis 138:1169–1179, 1988.
receiving mechanical ventilation for > 48 h. Chest 133:853–861, 92. Unroe M, MacIntyre NR: Evolving approaches to assessing and
2008. monitoring patient ventilatory interactions. Curr Opin Crit Care
68. Yilmaz M, Keegan MT, Iscimen R, et al: Toward the prevention of 16(3):261–268, 2010.
acute lung injury: protocol-guided limitation of large tidal volume 93. Yang LY, Huang YC, MacIntyre NR: Patient-ventilator synchrony
ventilation and inappropriate transfusion. Crit Care Med 35:1660– during pressure-targeted versus flow-targeted small tidal volume
1666, 2007. assisted ventilation. J Crit Care 22:252–257, 2007.
69. Villar J, Kacmarek RM, Perez-Mendez L, Aguirre-Jaime A: A high 94. Jubran A, Van de Graaf WB, Tobin MJ: Variability of patient ventila-
positive end-expiratory pressure, low tidal volume ventilatory strat- tor interactions with pressure support ventilation in patients with
egy improves outcome in persistent acute respiratory distress syn- chronic obstructive pulmonary disease. Am J Respir Crit Care Med
drome: a randomized, controlled trial. Crit Care Med 34:1311–1318, 152:129–136, 1995.
2006. 95. Epstein SK: How often does patient-ventilatory asynchrony
70. NIH ARDS Network: Ventilation with lower tidal volumes as com- occur and what are the consequences? Respir Care 56(1):25–38,
pared with traditional tidal volumes for acute lung injury and the 2011.
acute respiratory distress syndrome. N Engl J Med 342:1301–1308, 96. Pohlman MC, et al: Excessive tidal volume from breath stacking
2000. during lung-protective ventilation for acute lung injury. Crit Care
71. Vaporidi K, Voloudakis G, Priniannakis G, et al: Effects of respiratory Med 36(11):3019–3023, 2008.
rate on ventilator-induced lung injury at a constant PaCO2 in a 97. Tejerina E, Frutos-Vivar F, Restrepo MI, et al: Incidence, risk factors,
mouse model of normal lung. Crit Care Med 36:1277–1283, 2008. and outcome of ventilator-associated pneumonia. J Crit Care 21:56–
72. Rich BR, Reickert CA, Sawada S, et al: Effect of rate and inspiratory 65, 2006.
flow on ventilator induced lung injury. J Trauma 49:903–911, 98. Collard HR, Saint S, Matthay M, et al: Prevention of ventilator-
2000. associated pneumonia: an evidence-based systematic review. Ann
73. Crotti S, Mascheroni D, Caironi P, et al: Recruitment and derecruit- Intern Med 138:494–501, 2003.
ment during acute respiratory failure. Am J Respir Crit Care Med 99. Dodek P, Keenan S, Cook D, et al: Evidence-based clinical practice
164:131–140, 2001. guideline for the prevention of ventilator-associated pneumonia.
74. Rimensberger PC, Prisine G, Mullen BM, et al: Lung recruitment Ann Intern Med 141:305–313, 2004.
during small tidal volume ventilation allows minimal positive end 100. Nachtigall I, Tamarkin A, Tafelski S, et al: Impact of adherence to
expiratory pressure without augmenting lung injury. Crit Care Med standard operating procedures for pneumonia on outcome of inten-
27:1940–1945, 1999. sive care unit patients. Crit Care Med 37:159–166, 2009.
101  •  Mechanical Ventilation 1777.e3

101. Pinciroli R, Mietto C, Berra L: Respiratory therapy device modifica- 124. Hurford WE, Cheifetz IM: Respiratory controversies in the critical
tions to prevent ventilator-associated pneumonia. Curr Opin Infect care setting: should heliox be used for mechanically ventilated
Dis 26(2):175–183, 2013. patients? Respir Care 52:582–594, 2007.
102. Palmer LB, Smaldone GC, Chen JJ, et al: Aerosolized antibiotics and 125. ACCP/AARC/SCCM Task Force: Evidence based guidelines for
ventilator-associated tracheobronchitis in the intensive care unit. weaning and discontinuing mechanical ventilatory support. Chest
Crit Care Med 36:2008–2013, 2008. 120(Suppl 6), 2001. [Also in Respir Care 47:20–35, 2002].
103. Hickling KG, Walsh J, Henderson S, Jackson R: Low mortality rate in 126. Peñuelas O, Frutos-Vivar F, Fernández C, et al: Characteristics and
adult respiratory distress syndrome using low-volume, pressure- outcomes of ventilated patients according to time to liberation from
limited ventilation with permissive hypercapnia: a prospective study. mechanical ventilation. Am J Respir Crit Care Med 184(4):430–437,
Crit Care Med 22:1568–1578, 1994. 2011.
104. MacIntyre NR: Is there a best way to set the tidal volume for mechan- 127. Ferrer M: Non-invasive ventilation in the weaning process. Minerva
ical ventilatory support? Clin Chest Med 29:225–232, 2008. Anestesiol 74:311–314, 2008.
105. MacIntyre NR: Is there a best way to set positive end expiratory pres- 128. Nava S, Gregoretti C, Fanfulla F, et al: Noninvasive ventilation to
sure for mechanical ventilatory support in acute lung injury? Clin prevent respiratory failure after extubation in high-risk patients. Crit
Chest Med 29:233–240, 2008. Care Med 33(11):2465–2470, 2005.
106. Checkley W, Brower R, Korpak A, et al: Effects of a clinical trial on 129. Girard TD, Kress JP, Fuchs BD, et al: Efficacy and safety of a paired
mechanical ventilation practices in patients with acute lung injury. sedation and ventilator weaning protocol for mechanically ventilated
Am J Respir Crit Care Med 177:1215–1222, 2008. patients in intensive care (Awakening and Breathing Controlled
107. Esteban A, Ferguson ND, Meade MO, et al: Evolution of mechanical Trial): a randomized controlled trial. Lancet 371:126–134, 2008.
ventilation in response to clinical research. Am J Respir Crit Care Med 130. Mehta S, Burry L, Cook D, et al: Daily sedation interruption in
177:170–177, 2008. mechanically ventilated critically ill patients cared for with a seda-
108. Amato MB, Barbas CSV, Medievos DM, et al: Effect of a protective tion protocol: a randomized controlled trial. JAMA 308(19):1985–
ventilation strategy on mortality in ARDS. N Engl J Med 338:347– 1992, 2012.
354, 1998. 131. Froese AB: High frenquency oscillatory ventilation for ARDS: let’s
109. Cooke CR, Kahn JM, Watkins TR, et al: Cost effectiveness of imple- get it right this time. Crit Care Med 25:906–908, 1998.
menting low tidal volume ventilation in patients with acute lung 132. Chang HK: Mechanisms of gas transport during high frequency ven-
injury. Chest 136:79–88, 2009. tilation. J Appl Physiol 56:553–563, 1984.
110. Papazian L, Forel JM, Gacouin A, et al: Neuromuscular blockers in 133. Fessler HE, Hager DN, Brower RG: Feasibility of very high-frequency
early acute respiratory distress syndrome. N Engl J Med 363(12): ventilation in adults with acute respiratory distress syndrome. Crit
1107–1116, 2010. Care Med 36(4):1043–1048, 2008.
111. Habashi NM: Other approaches to open-lung ventilation: airway 134. Bollen CW, Uiterwal SPM, vanVught AJ: Cumulative meta analysis
pressure release ventilation. Crit Care Med 33(3 Suppl ):S228–S240, of high frequency vs conventional ventilation in premature neo-
2005. nates. Am J Respir Crit Care Med 168:1150–1155, 2003.
112. Putensen C, Zech S, Wrigge H, et al: Long term effects of spontane- 135. Courtney SE, Durand DJ, Asselin JM, et al: High frequency oscillatory
ous breathing during ventilatory support in patients with ALI. Am J ventilation vs conventional mechanical ventilation for very low birth
Respir Crit Care Med 164:43–49, 2001. weight infants. N Engl J Med 347:643–652, 2002.
113. Varpula T, Valta P, Niemi R, et al: Airway pressure release ventilation 136. Sud S, Sud M, Friedrich JO, et al: High frequency oscillation in
as a primary ventilatory mode in acute respiratory distress syn- patients with acute lung injury and acute respiratory distress syn-
drome. Acta Anaesth Scand 48:722–731, 2004. drome (ARDS): systematic review and meta-analysis. BMJ
114. Maxwell RA, Green JM, Waldrop J, et al: A randomized prospective 18(340):c2327, 2010.
trial of airway pressure release ventilation and low tidal volume 137. Young D, Lamb SE, Shah S, et al: High-frequency oscillation for
ventilation in adult trauma patients with acute respiratory failure. acute respiratory distress syndrome. N Engl J Med 368(9):806–813,
J Trauma 69(3):501–510, 2010. 2013.
115. González M, Arroliga AC, Frutos-Vivar F, et al: Airway pressure 138. Ferguson ND, Cook DJ, Guyatt GH, et al: High-frequency oscillation
release ventilation versus assist-control ventilation: a comparative in early acute respiratory distress syndrome. N Engl J Med 368(9):
propensity score and international cohort study. Intensive Care Med 795–805, 2013.
36(5):817–827, 2010. 139. Albert RK: Prone ventilation. Clin Chest Med 21(3):511–517, 2000.
116. Ranieri VM, Giuliani R, Fiore T, et al: Volume-pressure curve of the 140. Guérin C, Reignier J, Richard JC, et al: Prone positioning in severe
respiratory system predicts effects of PEEP in ARDS: “occlusion” acute respiratory distress syndrome. N Engl J Med 368(23):2159–
versus “constant flow” technique. Am J Respir Crit Care Med 149:19– 2168, 2013.
27, 1994. 141. Petter AH, Chiolero RL, Cassina T, et al: Automatic “respirator/
117. Suter PM, Fairley HB, Isenberg MD: Optimal end expiratory pressure weaning” with adaptive support ventilation: the effect on duration
in patients with acute pulmonary failure. N Engl J Med 292:284– of endotracheal intubation and patient management. Anesth Analg
289, 1975. 97:1743–1750, 2003.
118. Grasso S, Terragni P, Mascia L, et al: Airway pressure-time curve 142. Lellouche F, Mancebo J, Jolliet P, et al: A multicenter randomized trial
profile (stress index) detects tidal recruitment/hyperinflation in experi- of computer-driven protocolized weaning from mechanical ventila-
mental acute lung injury. Crit Care Med 32:1018–1027, 2004. tion. Am J Respir Crit Care Med 174:894–900, 2006.
119. Meade MO, Cook DJ, Guyatt GH, et al: Ventilation strategy using low 143. Rose L, Presneill JJ, Johnston L, Cade JF: A randomised, controlled
tidal volumes, recruitment maneuvers, and high positive end- trial of conventional versus automated weaning from mechanical
expiratory pressure for acute lung injury and acute respiratory dis- ventilation using SmartCare/PS. Intensive Care Med 34(10):1788–
tress syndrome: a randomized controlled trial. JAMA 299:637–645, 1795, 2008.
2008. 144. Hess DR, MacIntyre NR: Ventilator discontinuation: why are we still
120. Mercat A, Richard JC, Vielle B, et al: Positive end-expiratory pressure weaning? Am J Respir Crit Care Med 184:392–394, 2011.
setting in adults with acute lung injury and acute respiratory dis- 145. Fabry B, Zappe D, Guttman J: Breathing pattern and additional work
tress syndrome: a randomized controlled trial. JAMA 299:646–655, of breathing in spontaneously breathing patients with different ven-
2008. tilatory demand during inspiratory pressure support and automatic
121. Brower RG, Lanken PN, MacIntyre N, et al: Higher versus lower posi- tube compensation. Intensive Care Med 23:545–552, 1997.
tive end-expiratory pressures in patients with the acute respiratory 146. Guttmann J, Haberthür C, Mols G, Lichtwarck-Aschoff M: Auto-
distress syndrome. N Engl J Med 351:327–336, 2004. matic tube compensation (ATC). Minerva Anestesiol 68(5):369–377,
122. Briel M, Meade M, Mercat A, et al: Higher vs lower positive end- 2002.
expiratory pressure in patients with acute lung injury and acute 147. Ho L, MacIntyre NR: Effects of initial flow rate and breath termina-
respiratory distress syndrome: systematic review and meta-analysis. tion criteria on pressure support ventilation. Chest 99:134–138,
JAMA 303(9):865–873, 2010. 1991.
123. Talmor D, Sarge T, Malhotra A, et al: Mechanical ventilation guided 148. Tokioka H, Tanaka T, Ishizu T, et al: The effect of breath termination
by esophageal pressure in acute lung injury. N Engl J Med 359(20): criterion on breathing patterns and the work of breathing during
2095–2104, 2008. pressure support ventilation. Anesth Analg 92(1):161–165, 2001.
1777.e4 PART 3  •  Clinical Respiratory Medicine

149. Lellouche F, Brochard L: Advanced closed loops during mechanical versus proportional assist ventilation. Crit Care Med 35(4):1048–
ventilation (PAV, NAVA, ASV, SmartCare). Clin Anaesthesiol 23:81– 1054, 2007.
93, 2009. 156. Ranieri VM, Giuliani R, Mascia L, et al: Patient ventilator interaction
150. Younes M: Proportional assist ventilation, a new approach to venti- during acute hypercapnia; pressure-support vs. proportional-assist
latory support. Am Rev Respir Dis 145:114–120, 1992. ventilation. J Appl Physiol 81:426–436, 1996.
151. Kacmarek RM: Proportional assist ventilation and neurally adjusted 157. Sinderby C: Neurally adjusted ventilatory assist (NAVA). Minerva
ventilatory assist. Respir Care 56(2):140–148, 2011. Anestesiol 68:378–380, 2002.
152. Kacmarek RM, Villar J: When it comes to ventilation, noisy is better 158. Spahija J, de Marchie M, Albert M, et al: Patient-ventilatory interac-
than quiet and variability is healthier than constant! Crit Care Med tion during pressure support ventilation and neurally adjusted ven-
39(4):898–899, 2011. tilatory assist. Crit Care Med 38(2):518–526, 2010.
153. Sinderby C, Beck J: Proportional assist ventilation and neurally 159. Lecomte F, Brander L, Jalde F, et al: Physiological response to increas-
adjusted ventilatory assist–better approaches to patient ventilatory ing levels of neurally adjusted ventilatory assist (NAVA). Respir
synchrony? Clin Chest Med 29(2):329–342, 2008. Physiol Neurobiol 166(2):117–124, 2009.
154. Grasso S, Puntillo F, Mascia L, et al: Compensation for increase in 160. Brander L, Leong-Poi H, Beck J, et al: Titration and implementation
respiratory workload during mechanical ventilation. Pressure- of neurally adjusted ventilatory assist in critically ill patients. Chest
support versus proportional-assist ventilation. Am J Respir Crit Care 135(3):695–703, 2009.
Med 161(3 Pt 1):819–826, 2000. 161. Spahija J, de Marchie M, Albert M, et al: Patient-ventilator interac-
155. Bosma K, Ferreyra G, Ambrogio C, et al: Patient-ventilator interac- tion during pressure support ventilation and neurally adjusted ven-
tion and sleep in mechanically ventilated patients; pressure support tilatory assist. Crit Care Med 38(2):518–526, 2010.

You might also like